Next Article in Journal
Influence of Nanoclay on the Thermo-Mechanical Properties of Silicone Pressure-Sensitive Adhesives
Next Article in Special Issue
Hydrocarbon Lubricating Oils with Admixture of Ionic Liquid as Electrorheological Medium
Previous Article in Journal
Simulation of Inclusion Particle Motion Behavior under Interfacial Tension in Continuous Casting Mold
Previous Article in Special Issue
Thermal Decomposition, Low Temperature Phase Transitions and Vapor Pressure of Less Common Ionic Liquids Based on the Bis(trifuoromethanesulfonyl)imide Anion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights into the Effect of Lithium Doping on the Deep Eutectic Solvent Choline Chloride:Urea

by
Giselle de Araujo Lima e Souza
1,
Maria Enrica Di Pietro
1,*,
Franca Castiglione
1,
Valeria Vanoli
1 and
Andrea Mele
1,2,*
1
Department of Chemistry, Materials and Chemical Engineering “G. Natta”, Politecnico di Milano, Piazza L. da Vinci 32, 20133 Milan, Italy
2
CNR-SCITEC Istituto di Scienze e Tecnologie Chimiche, Via A. Corti 12, 20133 Milan, Italy
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(21), 7459; https://doi.org/10.3390/ma15217459
Submission received: 16 September 2022 / Revised: 17 October 2022 / Accepted: 21 October 2022 / Published: 24 October 2022
(This article belongs to the Special Issue The Wonderful World of Ionic Liquids and Deep Eutectic Solvents)

Abstract

:
Choline-based deep eutectic solvents (DESs) are potential candidates to replace flammable organic solvent electrolytes in lithium-ion batteries (LIBs). The effect of the addition of a lithium salt on the structure and dynamics of the material needs to be clarified before it enters the battery. Here, the archetypical DES choline chloride:urea at 1:2 mole fraction has been added with lithium chloride at two different concentrations and the effect of the additional cation has been evaluated with respect to the non-doped system via multinuclear NMR techniques. 1H and 7Li spin-lattice relaxation times and diffusion coefficients have been measured between 298 K and 373 K and revealed a decrease in both rotational and translational mobility of the species after LiCl doping at a given temperature. Temperature dependent 35Cl linewidths reflect the viscosity increase upon LiCl addition, yet keep track of the lithium complexation. Quantitative indicators such as correlation times and activation energies give indirect insights into the intermolecular interactions of the mixtures, while lithium single-jump distance and transference number shed light into the lithium transport, being then of help in the design of future DES electrolytes.

1. Introduction

The ever-increasing global energy demand is constantly asking for the development of new safe, cheap and environment-friendly energy storage devices. Among the electrochemical energy storage systems, lithium ion batteries (LIBs) represent one of the most successful technologies with a dominant influence on the modern society in a plethora of applications from portable electronic devices, to medical technology, electric and hybrid vehicles, and stationary storage [1,2,3]. However, when subjected to thermal, mechanical and electrical abuses, LIBs can abruptly release chemical energy in the form of fire or explosions, and possible leakage of hazardous chemicals, leading to disastrous accidents involving different devices from cell phones and laptops, to electric vehicles and airplanes [1]. In such uncontrollable thermal runaway processes, liquid electrolytes play a significant role. Currently, state-of-the-art electrolytes in LIBs are composed of a thermally unstable mostly fluorinated lithium salt (usually lithium hexafluorophosphate, LiPF6) dissolved in highly flammable organic solvents such as ethylene carbonate, dimethyl carbonate, propylene carbonate and diethylcarbonate [1,2,3]. These electrolytes act as the “fuel” for the battery combustion and explosion in case of a thermal runaway process, and cause in general adverse environmental impacts at the end of the life of the batteries [1,3]. Such concerns over the safety hazard, waste management and environmental impact of LIBs are currently driving the academic and industrial communities to consider alternative non-flammable and environmentally friendly solvents as replacement of the organic electrolytes.
Deep Eutectic Systems (DESs) stand out as a promising option in this respect [4,5]. They inherit the appreciable ionic conductivity and low- or non-flammability of the similar compounds room temperature ionic liquids (ILs) [6], but do not suffer from the same downsides, i.e., high cost, quite complicated synthesis and questioned safety. On the contrary, DESs are endowed with additional advantages such as air and moisture stability, higher solubility for metal salts and wider tunability of functional properties. Besides, DESs are composed of largely available, environmentally friendly and safe raw materials, and they are also easy to prepare with a 100% atom economy and no need for purification steps [4,5]. The cost-efficiency and high sustainability of DESs make them extremely interesting as benign-by-design electrolyte materials in energy storage devices, and promising candidates to replace organic solvents in LIBs.
Although a strict and widely recognized definition does not exist at present, DESs usually consist of a mixture of at least two components, normally solid at room temperature, which self-associate when mixed at a given mole ratio, with formation of a “deep” eutectic mixture. The establishment of intermolecular interactions, mainly–but not only–hydrogen bonds (H-bonds), results in a marked depression of the melting point around the eutectic point with respect to both the pure constituents and the one predicted by assuming a thermodynamic ideal behavior of the solid-liquid equilibrium [4,7].
Among the different families of DESs (Table S1) [8,9], type IV DESs (mixtures of a metal salt and a molecular hydrogen bond donor, HBD) dominate the electrochemical field, while only a few applications of type III DESs (composed of an organic salt and a molecular HBD) are reported [5,10,11]. The claimed advantage of the former (and vice versa drawback of the latter) would be the presence of Li+ as single cation species in the mixture, thus enabling the achievement of higher transference numbers [11,12,13]. It should be noted, however, that preparing those type IV DESs requires large amounts of expensive lithium salt, rising the production costs when compared to dissolving controlled amounts of the same salt in low-cost type III DESs, hence jeopardizing one of the merits of DESs. In a previous work [14], we studied a prototypical system composed of choline chloride:urea at 1:2 mole fraction (ChCl:U xChCl = 0.33) with LiCl dissolved at a concentration of 0.4 mol/kg. A combination of multinuclear NMR measurements and molecular dynamics simulations pointed towards a strong coordination of Li+ cations by chloride anions and a marked involvement of U in the H-bond network. These preliminary results suggest a systematic evaluation of the effect of lithium doping in the underexploited type III DESs electrolyte solutions. Actually, the addition of a small positively charged ion, Li+, is known to affect both the macroscopic (e.g., viscosity [15]) and the microscopic (e.g., molecular diffusion [16]) properties of ChCl. Since the peculiar macro- and microscopic behavior of DESs stems from a tangled–and often not trivially predictable–balance of intermolecular forces, this calls for a profound understanding, at molecular /atomic level, of the effect of the addition of lithium ions in such network, for the design, optimization, and development of deep eutectic electrolytes (DEEs).
In the present work, we contribute to the experimental design mentioned above by increasing lithium contents (0.8 and 1.0 mol/kg) and comparing the Li-doped systems with the pristine Li-free ChCl:U. This allows to gather more insights into the effect of Li-doping and Li-concentration on the dynamic behavior of the different species in the mixtures. 35Cl NMR spectra were collected over a 75 K temperature interval in 5 K temperature steps, with the linewidths reflecting both the viscosity increase after LiCl addition and the underlying Li-Cl complexation. 1H and 7Li spin-lattice relaxation and diffusion measurements were performed as a function of temperature, and interpreted in terms of correlation times, rotational and translational activation energies, transference number and lithium single-jump distance.

2. Materials and Methods

All materials, choline chloride (ChCl, ≥ 98%), urea (U, ≥ 98%) and lithium chloride (LiCl, ≥99%) were purchased from Sigma Aldrich and vacuum dried prior to use. The DES ChCl:U at 1:2 molar ratio (xChCl = 0.33) was prepared by mixing the two components, ChCl and U, at the proper mole fractions and heating at 80 °C under constant stirring until the formation of a homogeneous liquid. Next to the pure DES (LiCl 0 wt%), two lithium-doped samples were prepared by adding LiCl to the mixture under stirring at two different concentrations (ca. 3 wt% and 4 wt%). The water content of the samples was below 2 wt% as measured by Karl-Fischer titration. A summary of the samples is given in Table 1, while the structure of the components is shown in Figure S1.
The samples were transferred to 5 mm NMR tubes, equipped with a capillary containing deuterated dimethylsulfoxide (DMSO-d6). NMR measurements were performed without sample spinning with a Bruker NEO 500 console (11.74 T) equipped with a direct observe BBFO (broadband including fluorine) iProbe and a variable-temperature unit (1H, 7Li and 35Cl resonance frequency of 500.13, 194.37 and 49.00 MHz, respectively). The instrument was carefully tuned, shimmed, and the 90° pulses calibrated.
1D 1H, 7Li and 35Cl, as well as 1H and 7Li T1 relaxation and diffusion experiments were performed over a temperature range of 298 K to 373 K, in 5 K increments, with a minimum of 15 min allowed for thermal equilibration. 1D 1H and 7Li spectra were recorded with 1 scan using 16384 points, over a spectral width of 9 and 50 ppm, respectively. 35Cl spectra were collected with standard 1D acquisition sequences, using the ARING pulse train to reduce acoustic ringing artefacts in 35Cl measurements. Automated peak-picking and linewidth routines were used for estimations of 35Cl full width at half maximum (FWHM).
1H and 7Li T1 relaxation measurements were carried out with the inversion recovery (IR) pulse sequence. Spectra were recorded using relaxation delays at least five times the longest T1, four dummy scans prior to acquisition, and data matrices of 16,384 (t2) × 16 (t1), with 2 transients per increment, over a spectral width of 9 ppm for 1H and 50 ppm for 7Li. The spin−lattice relaxation rates were measured for delay times ranging from 0.05–5 s to 0.05–10 s, according to the temperature. The baselines of all arrayed T1 spectra were corrected prior to processing the data using an exponential filter in F2 dimension (with LB equal to 0.3 Hz). Relaxation times were computed using integrals from experimental raw data by means of the Bruker T1/T2 relaxation module using the manual integration option and applying the standard one-component fitting function. Fits to extract the correlation times and rotational activation energies from the relaxation rates were performed with OriginPro 2018 using a user-defined function with the Levenberg–Marquardt algorithm.
1H and 7Li self-diffusion coefficients were measured by pulsed field gradient (PFG)-NMR experiments by applying sine-shaped pulsed magnetic field gradients along the z-direction up to a maximum strength of G = 48.15 G cm−1. The diffusion experiments were performed using the bipolar pulse longitudinal eddy current delay (BPP-LED) pulse sequence. All experiments were carried out using 16384 points in the F2 dimension, over a spectral width of 9 ppm for 1H and 50 ppm for 7Li, with a total of eight transients per increment. The relaxation delay was set to at least five times T1, and four dummy scans were programmed prior to acquisition. The pulse gradients were incremented from 2 to 95% of the maximum gradient strength in a linear ramp with 32 steps. For each DOSY experiment, the duration of the magnetic field pulse gradients (δ) and the diffusion times (Δ) were optimized to obtain, where possible, 95% signal attenuation for the slowest diffusion species in the last step experiment. For 1H experiments, δ values were in the 2.6–6.0 ms range, while Δ values were 0.3–0.8 s long. For 7Li experiments, δ and Δ values were equal to 6.0 ms and 0.8 s, respectively. Due to slow motion, not enough signal attenuation was achieved at the lowest temperatures, then diffusion data for lithium were collected in the reduced temperature range 318–373 K. The baselines of all arrayed spectra were corrected prior to processing the data. Data were processed using an exponential filter in F2 dimension (LB = 0.3 Hz), and integrals were used in calculating relaxation times. The Stejskal-Tanner equation [17] was used to extract the self-diffusion coefficients for each peak via the Bruker T1/T2 module of TopSpin. Linear fits of the Arrhenius plots were performed with OriginPro 2018.

3. Results and Discussion

At the light of the results obtained on archetypal lithium-doped ChCl:U [14], here we propose a more systematic investigation on two samples containing a higher amount of metal salt (0.8 and 1.0 mol/kg), and compare the results with the pure DES. This allows to gather novel insights into the effect of Li-doping and Li-concentration on the dynamic behavior of the different species in the mixtures. Note that ChCl and most type III DESs are highly hygroscopic, hence preventing the preparation of a water-free system when working under ambient conditions. We found previously that a small amount of water (3.9 wt%) substantially does not alter the intermolecular structure of the ChCl DES electrolyte, but has a remarkable effect on dynamics [14]. This is in line with the by now established concept that a controlled amount of water does not seriously break up the intermolecular network of neat DESs [18], but has a non-negligible effect in reducing viscosity. Due to preliminary vacuum drying, a residual amount of water (2 wt%) must be considered here for the three samples, which is close to a realistic scenario for industrial processes carried out at ambient air conditions.
Although, when designing a DEE, the obvious interest is focused on lithium, the study of its local environment and mobility cannot leave aside all other components of the mixtures. The NMR approach applied in the present work allows from the one hand to probe individually the rotational and translational mobility of the different species and on the other to gain an overall picture of the intermolecular network within the mixtures. In this way, we gain a deeper understanding of how the strength of the intermolecular interactions in the DEE change at increasing lithium concentration and in comparison to the Li-free system. We discuss in the following the results obtained from chemical shift/linewidth analysis, relaxation and diffusion measurements on the different species in the mixtures, with a special focus on quantitative descriptors of lithium transport.

3.1. 1D 1H, 7Li and 35Cl NMR: Temperature-Dependent Chemical Shift and Linewidth

The analysis of chemical shift variations gives preliminary hints into the strength of the intermolecular interactions experienced by the different species (Figure 1). As expected, an upfield shift is observed for all sites, indicating a weakening of the interactions due to the increase in temperature. As a result of their major involvement in the H-bond network, the peaks of U and OH are the most sensitive to the temperature change. Choline signals are barely affected, while Li shows a moderate upfield shift. Noteworthy is that the addition of Li salt does not have a drastic impact on the temperature-dependent 1H chemical shifts (for instance, | δ ( 373   K ) δ ( 298   K ) | = | Δ δ m a x | is in the range 0.3–0.5 ppm for U and OH, respectively, in all samples). The presence of LiCl is instead affecting the linewidth of 35Cl signal, causing a significant increase of the full width at half maximum (FWHM) with respect to the neat DES (Figure S2), which is an expected consequence of the viscosity increase. As expected, the increase in temperature causes a marked drop in linewidth of the 35Cl signal, likely due to the decrease in viscosity. Perusal of Figure S2b also highlights a change in slope of the curve of 35Cl NMR linewidth vs T at about 335 K for the two Li containing samples, while no evident change is observed in the curve related to the DES without LiCl. Thus, the profile of the temperature-dependent FWHMs in the presence of LiCl shows a smaller gradient in the first T region (up to roughly 335 K) with respect to the non-doped system. Given the 35Cl linewidth is indicative of the Cl- local environment [19], this would qualitatively indicate that, in the presence of Li+, the chloride local environment is less sensitive to the temperature increase–reasonably due to Li complexation. Data for ChCl:U-LiCl(0.8) and ChCl:U-LiCl(1.0) mostly overlap, thus suggesting that both the viscosity and the symmetry of the chloride environment are comparable in the two lithium-doped systems.

3.2. 1H and 7Li PFG-NMR: Self-Diffusion Coefficients, Translational Activation Energy and Transference Number

PFG-NMR was applied to collect the self-diffusion coefficients of the different species in the mixtures over a 75 K temperature interval, using 5 K steps.
Table S2 and Figure 2a show the self-diffusion coefficients measured for the different species in the three samples by PFG-NMR. Diffusivity follows the order U > Ch > Li. Lithium is hence the slowest species in the mixtures, due to the strong interactions involving this cation. Indeed, Li+ interacts strongly with chloride anions, as already observed for the parent system with LiCl 0.4 mol/kg [14] and is widely known for similar systems like ionic liquid mixtures [20,21]. An additional contribution to the slow translational dynamics of Li+ stems from its interactions with urea in the second coordination shell [14]. The addition of LiCl 0.8 mol/kg roughly halves the diffusion of both U and Ch, which is further slowed down when more LiCl is added. This is intuitively related to the increased viscosity of the system once the metal salt is added.
The apparent activation energies for the translational motion, E a t r a n s l , were obtained fitting the diffusion data to an Arrhenius expression and are reported in Figure 2b and Table S3. In the two Li-doped samples, E a t r a n s l is in the order: Li U > Ch. The gap in activation energy between Li and Ch seems to be inversely proportional to the LiCl content (2.9 kJ/mol < 5.1 kJ/mol < ca. 7 kJ/mol for concentrations of 1.0 mol/kg, 0.8 mol/kg and 0.4 mol/kg, respectively) [14]. In the previous work [14], we highlighted a peculiar behavior of U, which is showing a diffusion behavior–and consequently a E a t r a n s l –strongly dependent on its average H-bond interactions in the system, and then on the temperature. In particular, the activation energy of U was found to be well correlated with Li–Cl complexes in the low T regime, and with Ch in the high T regime. Here we observe that the non-Arrhenius behavior of urea diffusivity is even more marked in the pure DES (R2 = 0.990), but, on the other hand, it is less relevant in the Li-doped systems (R2 = 0.996 and 0.998 for LiCl 0.8 mol/kg and 1.0 mol/kg, respectively). Increased amount of LiCl would reduce the temperature-dependence of U activation energy, making it closer to that of lithium itself in all the investigated temperature range (53.4 vs. 53.2 kJ/mol in ChCl:U-LiCl(0.8) and 55.11 vs. 54.7 kJ/mol in ChCl:U-LiCl(1.0)). Bearing in mind that Li interacts with urea in the second coordination shell [14], this would indicate a strengthening of such interactions with increased Li content.
A convenient way to quantify the relative mobility of Li ions with respect to other species in the mixture is via the transference number, which is the fractional contribution of an ion species to the overall conductivity [22]. Indeed, all ions in the system contribute to the total current, but Li ion transport is crucial for the charge/discharge process of LIBs [23]. Under the assumption that all ionic species are free, i.e., ion pairs and/or ion clustering are neglected, an apparent transference number for Li can be calculated from NMR diffusion experiments as given by t L i + = n L i + D L i + / i n i D i , with n i the number of charge carriers [23]. Given the diffusion coefficient of the chloride anion is not directly accessible by PFG-NMR, a Li ion transference number cannot be calculated. However, previous MD simulations in ChCl:U systems found that chloride diffusivity is intermediate between Ch and U [24], with values closer to Ch in the presence of the additional Li cation [14]. Hence, we can define two boundary conditions for t L i + by considering D C l equal to the self-diffusion measured for the fastest (U) or the slowest (Li) species (Figure 3), respectively. This gives us the opportunity to estimate the upper and lower transference number of Li ion in this new DES-based electrolyte and compare with other DESs and DES analogs. As observed in Figure 3, estimated t L i + ranges for the two samples from 0.04 to 0.09, with a negligible change when increasing the LiCl content. These values are much smaller than those found for type IV DESs composed of lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) and urea or a sulfonamide as HBD ( t L i + between 0.4 and 0.7 in the temperature range 298–353 K) [13,25]. t L i + values between 0.44 and 0.61 at 30 °C were also reported by Watanabe and coworkers in solvated IL systems consisting of equimolar mixtures of a lithium salt and a glyme [12,26]. However, values found here for ChCl:U-LiCl are in line or even higher than transference numbers found for traditional mixtures of ILs and lithium salts (i.e., t L i + in the range 0.02–0.15 are common values for alkyl-imidazolium and alkyl-pyrrolidinium ILs doped with lithium salts [27,28,29]). Therefore, even bearing in mind that apparent transference numbers are only an approximate concept that does not take into account correlated motion of ions, the values found here for the prototypical ChCl:U-LiCl system are probably too low for battery applications at present. The high tunability of DES offers, however, large room for improvement and justify–or even push–further investigations on developing other DEEs.

3.3. 1H and 7Li Relaxation NMR: Correlation Times, Rotational Activation Energy and Lithium Single-Jump Distance

1H and 7Li T1 relaxation times were measured between 298 K to 373 K (Figure S3). Curves of Ch protons pass through a minimum at lower temperature than U protons, while Li curves approach a minimum without reaching it. This is a first indication that Ch protons experience faster local dynamics compared to U and Li. To quantify the rotational motion at each site/species, the Bloembergen, Purcell, and Pound (BPP) model [14,21,30,31,32,33,34,35] was applied to the spin-lattice temperature-dependent relaxation data. The obtained correlation time τ C and apparent activation energy for the rotational motion E a r o t are reported in Figure 4 and Tables S4–S7. As already observed [14], the T1 curve of U is not reproduced by the BPP model at low temperature, and the divergence gets more pronounced as the LiCl content increases. A good fit of T1 data for urea could be obtained in the interval 308–373 K, 323–373 K and 333–373 K for samples ChCl:U-LiCl(0), ChCl:U-LiCl(0.8) and ChCl:U-LiCl(1.0), respectively. Note, however, that no significant differences are observed in the output parameters when fitting data for ChCl:U-LiCl(0) and ChCl:U-LiCl(0.8) in the reduced interval used for ChCl:U-LiCl(1.0) (Table S4).
Applying the BPP equation for the dominant 1H-1H dipolar relaxation mechanism, the correlation time of the dipolar interaction τ C ( H ) can be obtained [21,31,32,36]. τ C ( H ) can be thought of as the time for a molecular rotation by roughly 1 radian about any axis [37]. Note that for small rigid molecules, it is representative of the whole molecular reorientational time constant [38], whereas for flexible molecules it contains also contributions from internal motions of each given segment [31,32,39]. In both Li-free and Li-doped systems, Ch protons have shorter τ C ( H ) –hence faster rotational motion–than U protons, in agreement with previous observations [14,31] and with the considerations drawn before. The addition of Li salt 0.8 mol/kg causes an increase in all τ C ( H ) compared to the neat DES (see also Figure S4 in the ESI). For instance at 333 K, τ C ( H ) values of Ch aliphatic protons slightly increase from 95–100 ps in ChCl:U-LiCl(0) to 115–125 ps in ChCl:U-LiCl(0.8), and τ C ( H ) values of U protons from 0.19 ns in ChCl:U-LiCl(0) to 0.30 ns in ChCl:U-LiCl(0.8). This indicates an overall slowdown of the rotational motion of all sites, which is compatible with a macroscopic increase of viscosity of the system. A further addition of LiCl does not have a marked effect, but it seems to affect differently urea and cholinium: while τ C ( H ) values of U protons slightly increase in ChCl:U-LiCl(1.0) (0.34 ns at 333 K), τ C ( H ) values of Ch protons slightly decrease (112–120 ps at 333 K). Even if interpreted with care, this different behavior would support again a scenario where interactions with U are strengthened on going from 0.8 mol/kg to 1.0 mol/kg of LiCl, while Ch protons are not strongly involved in the network.
To give a physical interpretation to τ C ( L i ) , it has to be considered that the dominant relaxation mechanism is represented here by the quadrupolar interaction. Such interaction arises from the coupling of the electric quadrupole moment –a property of the nucleus– with an electric field gradient (EFG), which is intrinsic to the sample [40]. As τ C ( L i ) reflects the fluctuations of the EFG, and given that the EFG is generated by the surroundings, a reasonable interpretation of τ C ( L i ) is less intuitive than τ C ( H ) and has to take into account both the local environment of the ion and the deformation of the electronic cloud of the ion itself induced by the surroundings [40]. A simplified model assumes that the first solvation shell plays the dominant role in the fluctuations of the EFG at the nucleus site and, in the specific case of the Li, oscillations of the EFG correspond to the moving of this ion in its solvation cage [41]. In this description, 7Li relaxation would occur by the quadrupolar mechanism for translational motion rather than rotational motion of lithium atom [34,35], and τ C ( L i ) would represent a single lithium jump from one position to another. This allows one to calculate an average distance for a Li one-jump as R o n e f l i p = 6 D τ C ( L i ) [34,35]. The estimated average jump distances for the two Li-doped systems investigated here are displayed in Figure 5. R o n e f l i p values range between 0.10 and 0.35 nm in the temperature window 318–373 K, hence slightly smaller than what previously calculated for the more hydrated parent system comparing the same temperature range [14]. However, the overall similarity of the trends would suggest an analogous Li transport within the mixture, i.e., lithium ions migrating via jumps within the solvation sphere.

4. Conclusions

Deep eutectic electrolytes show high attractiveness and competitiveness with respect to organic electrolytes and ILs and hold great promise for LIB applications. However, compared to the former cases, which are substantially mixtures of a solvent and a lithium salt, a rationalization of Li+ solvation and transport is even more difficult to achieve in the case of DES-based Li electrolytes, which are characterized by a higher system complexity due to both the higher number of actors and the variety of intermolecular interactions among them. Urea’s dynamic behavior emerged as paradigmatic: diffusion data showed that the participation of U to the network of interactions is not only temperature-dependent but also LiCl concentration-dependent. This adds fundamental knowledge to the type of intermolecular interactions taking place in the solvation shells and opens the possibility of modulating the transport properties of the system. In general, the prototypical systems ChCl:U-LiCl studied here rise interest not only for possible applications, but also as probes for exploring the effect of Li-doping on the structure and dynamics of the different species in the mixtures. Multinuclear NMR allows to monitor–at atomic level–the dynamics of the single components. Additionally, the different dominant mechanisms of nuclear relaxation (dipolar and quadrupolar) previously discussed allow to draw a picture of the rotational and translational dynamics of the single components. In the case of 7Li relaxation, the dominant “jump” mechanism of diffusion outlined here provides fundamental information for DEE design. All considered, the picture emerging from this study shows high tunability of DEEs and exceptional dissolution abilities, which are the elements to overcome the low mobility of the Li+ cation found, a potential showstopper for future applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15217459/s1, Figure S1: Chemical structures; Figure S2: 35Cl NMR full width at half maximum; Figure S3: 1H and 7Li T1 relaxation times; Figure S4: 1H correlation time. Table S1: The main types of DESs; Table S2: Temperature-dependent 1H and 7Li self-diffusion coefficients; Table S3: Parameters of the linear fit of the Arrhenius plot of the diffusion data; Table S4: Best-fit parameters of the BPP model; Tables S5–S7: 1H and 7Li rotational correlation times. Reference [8,9] are cited in Supplementary Materials.

Author Contributions

Conceptualization, M.E.D.P.; methodology, M.E.D.P.; formal analysis, M.E.D.P.; investigation, M.E.D.P. and G.d.A.L.e.S.; resources, A.M.; data curation, M.E.D.P. and G.d.A.L.e.S.; writing—original draft preparation, M.E.D.P. and G.d.A.L.e.S.; writing—review and editing, M.E.D.P., G.d.A.L.e.S., F.C., V.V. and A.M.; visualization, M.E.D.P. and G.d.A.L.e.S.; supervision, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

A.M. thanks the Regione Lombardia and ENEA for the grant for G.A.L.S. Ph.D. programme. M.E.D.P. thanks Politecnico di Milano for her postdoctoral fellowship in the framework of the “MSCA EF Master Class 2018” funding programme.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, X.; Sun, Q.; Zhen, C.; Niu, Y.; Han, Y.; Zeng, G.; Chen, D.; Feng, C.; Chen, N.; Lv, W.; et al. Recent Progress in Flame-Retardant Separators for Safe Lithium-Ion Batteries. Energy Storage Mater. 2021, 37, 628–647. [Google Scholar] [CrossRef]
  2. Armand, M.; Axmann, P.; Bresser, D.; Copley, M.; Edström, K.; Ekberg, C.; Guyomard, D.; Lestriez, B.; Novák, P.; Petranikova, M.; et al. Lithium-Ion Batteries—Current State of the Art and Anticipated Developments. J. Power Sources 2020, 479, 228708. [Google Scholar] [CrossRef]
  3. Yang, Y.; Okonkwo, E.G.; Huang, G.; Xu, S.; Sun, W.; He, Y. On the Sustainability of Lithium Ion Battery Industry—A Review and Perspective. Energy Storage Mater. 2021, 36, 186–212. [Google Scholar] [CrossRef]
  4. Hansen, B.B.; Spittle, S.; Chen, B.; Poe, D.; Zhang, Y.; Klein, J.M.; Horton, A.; Adhikari, L.; Zelovich, T.; Doherty, B.W.; et al. Deep Eutectic Solvents: A Review of Fundamentals and Applications. Chem. Rev. 2021, 121, 1232–1285. [Google Scholar] [CrossRef]
  5. Di Pietro, M.E.; Mele, A. Deep Eutectics and Analogues as Electrolytes in Batteries. J. Mol. Liq. 2021, 338, 116597. [Google Scholar] [CrossRef]
  6. Watanabe, M.; Thomas, M.L.; Zhang, S.; Ueno, K.; Yasuda, T.; Dokko, K. Application of Ionic Liquids to Energy Storage and Conversion Materials and Devices. Chem. Rev. 2017, 117, 7190–7239. [Google Scholar] [CrossRef] [Green Version]
  7. Martins, M.A.R.; Pinho, S.P.; Coutinho, J.A.P. Insights into the Nature of Eutectic and Deep Eutectic Mixtures. J. Solution Chem. 2019, 48, 962–982. [Google Scholar] [CrossRef] [Green Version]
  8. Smith, E.L.; Abbott, A.P.; Ryder, K.S. Deep Eutectic Solvents (DESs) and Their Applications. Chem. Rev. 2014, 114, 11060–11082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Abranches, D.O.; Martins, M.A.R.; Silva, L.P.; Schaeffer, N.; Pinho, S.P.; Coutinho, J.A.P. Phenolic Hydrogen Bond Donors in the Formation of Non-Ionic Deep Eutectic Solvents: The Quest for Type v Des. Chem. Commun. 2019, 55, 10253–10256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Lv, Y.; Xiao, Y.; Ma, L.; Zhi, C.; Chen, S. Recent Advances in Electrolytes for “Beyond Aqueous” Zinc-Ion Batteries. Adv. Mater. 2022, 34, 2106409. [Google Scholar] [CrossRef]
  11. Lu, X.; Hansen, E.J.; He, G.; Liu, J. Eutectic Electrolytes Chemistry for Rechargeable Zn Batteries. Small 2022, 18, 2200550. [Google Scholar] [CrossRef] [PubMed]
  12. Ueno, K.; Yoshida, K.; Tsuchiya, M.; Tachikawa, N.; Dokko, K.; Watanabe, M. Glyme-Lithium Salt Equimolar Molten Mixtures: Concentrated Solutions or Solvate Ionic Liquids? J. Phys. Chem. B 2012, 116, 11323–11331. [Google Scholar] [CrossRef] [PubMed]
  13. Pauric, A.D.; Halalay, I.C.; Goward, G.R. Combined NMR and Molecular Dynamics Modeling Study of Transport Properties in Sulfonamide Based Deep Eutectic Lithium Electrolytes: LiTFSI Based Binary Systems. Phys. Chem. Chem. Phys. 2016, 18, 6657–6667. [Google Scholar] [CrossRef] [PubMed]
  14. Di Pietro, M.E.; Goloviznina, K.; van den Bruinhorst, A.; de Araujo Lima e Souza, G.; Costa Gomes, M.; Padua, A.A.H.; Mele, A. Lithium Salt Effects on the Liquid Structure of Choline Chloride—Urea Deep Eutectic Solvent. ACS Sustain. Chem. Eng. 2022, 10, 11835–11845. [Google Scholar] [CrossRef]
  15. Dhingra, D.; Bhawna; Pandey, S. Effect of Lithium Chloride on the Density and Dynamic Viscosity of Choline Chloride/Urea Deep Eutectic Solvent in the Temperature Range (303.15–358.15) K. J. Chem. Thermodyn. 2019, 130, 166–172. [Google Scholar] [CrossRef]
  16. Barik, S.; Chakraborty, M.; Sarkar, M. How Does Addition of Lithium Salt Influence the Structure and Dynamics of Choline Chloride-Based Deep Eutectic Solvents? J. Phys. Chem. B 2020, 124, 2864–2878. [Google Scholar] [CrossRef]
  17. Stejskal, E.O.; Tanner, J.E. Spin Diffusion Measurements: Spin Echoes in the Presence of a Time-Dependent Field Gradient. J. Chem. Phys. 1965, 42, 288–292. [Google Scholar] [CrossRef] [Green Version]
  18. Hammond, O.S.; Bowron, D.T.; Edler, K.J. The Effect of Water upon Deep Eutectic Solvent Nanostructure: An Unusual Transition from Ionic Mixture to Aqueous Solution. Angew. Chem.-Int. Ed. 2017, 56, 9782–9785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Di Pietro, M.E.; Hammond, O.; van den Bruinhorst, A.; Mannu, A.; Padua, A.; Mele, A.; Costa Gomes, M. Connecting Chloride Solvation with Hydration in Deep Eutectic Systems. Phys. Chem. Chem. Phys. 2021, 23, 107–111. [Google Scholar] [CrossRef]
  20. Bolimowska, E.; Castiglione, F.; Devemy, J.; Rouault, H.; Mele, A.; Pádua, A.A.H.; Santini, C.C. Investigation of Li+ Cation Coordination and Transportation, by Molecular Modeling and NMR Studies, in a LiNTf2-Doped Ionic Liquid-Vinylene Carbonate Mixture. J. Phys. Chem. B 2018, 122, 8560–8569. [Google Scholar] [CrossRef] [PubMed]
  21. Nürnberg, P.; Lozinskaya, E.I.; Shaplov, A.S.; Schönhoff, M. Li Coordination of a Novel Asymmetric Anion in Ionic Liquid-in-Li Salt Electrolytes. J. Phys. Chem. B 2020, 124, 861–870. [Google Scholar] [CrossRef] [PubMed]
  22. Gouverneur, M.; Kopp, J.; Van Wüllen, L.; Schönhoff, M. Direct Determination of Ionic Transference Numbers in Ionic Liquids by Electrophoretic NMR. Phys. Chem. Chem. Phys. 2015, 17, 30680–30686. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ye, H.; Huang, J.; Xu, J.J.; Khalfan, A.; Greenbaum, S.G. Li Ion Conducting Polymer Gel Electrolytes Based on Ionic Liquid/PVDF-HFP Blends. J. Electrochem. Soc. 2007, 154, A1048. [Google Scholar] [CrossRef] [Green Version]
  24. Sapir, L.; Harries, D. Restructuring a Deep Eutectic Solvent by Water: The Nanostructure of Hydrated Choline Chloride/Urea. J. Chem. Theory Comput. 2020, 16, 3335–3342. [Google Scholar] [CrossRef] [PubMed]
  25. Nandy, A.; Smiatek, J. Mixtures of LiTFSI and Urea: Ideal Thermodynamic Behavior as Key to the Formation of Deep Eutectic Solvents? Phys. Chem. Chem. Phys. 2019, 21, 12279–12287. [Google Scholar] [CrossRef] [PubMed]
  26. Tamura, T.; Hachida, T.; Yoshida, K.; Tachikawa, N.; Dokko, K.; Watanabe, M. New Glyme-Cyclic Imide Lithium Salt Complexes as Thermally Stable Electrolytes for Lithium Batteries. J. Power Sources 2010, 195, 6095–6100. [Google Scholar] [CrossRef]
  27. Lesch, V.; Jeremias, S.; Moretti, A.; Passerini, S.; Heuer, A.; Borodin, O. A Combined Theoretical and Experimental Study of the Influence of Different Anion Ratios on Lithium Ion Dynamics in Ionic Liquids. J. Phys. Chem. B 2014, 118, 7367–7375. [Google Scholar] [CrossRef] [Green Version]
  28. Frömling, T.; Kunze, M.; Schönhoff, M.; Sundermeyer, J.; Roling, B. Enhanced Lithium Transference Numbers in Ionic Liquid Electrolytes. J. Phys. Chem. B 2008, 112, 12985–12990. [Google Scholar] [CrossRef]
  29. Liu, X.; Mariani, A.; Diemant, T.; Di Pietro, M.E.; Dong, X.; Kuenzel, M.; Mele, A.; Passerini, S. Difluorobenzene-Based Locally Concentrated Ionic Liquid Electrolyte Enabling Stable Cycling of Lithium Metal Batteries with Nickel-Rich Cathode. Adv. Energy Mater. 2022, 12, 2200862. [Google Scholar] [CrossRef]
  30. Bloembergen, N.; Purcell, E.M.; Pound, R.V. Relaxation Effects in Nuclear Magnetic Resonance Absorption. Phys. Rev. 1948, 73, 679–712. [Google Scholar] [CrossRef] [Green Version]
  31. Triolo, A.; Di Pietro, M.E.; Mele, A.; Lo Celso, F.; Brehn, M.; Di Lisio, V.; Martinelli, A.; Chater, P.; Russina, O. Liquid Structure and Dynamics in the Choline Acetate:Urea 1:2 Deep Eutectic Solvent. J. Chem. Phys. 2021, 154, 244501. [Google Scholar] [CrossRef] [PubMed]
  32. Di Pietro, M.E.; Castiglione, F.; Mele, A. Polar/Apolar Domains ’ Dynamics in Alkylimidazolium Ionic Liquids Unveiled by the Dual Receiver NMR 1 H and 19 F Relaxation Experiment. J. Mol. Liq. 2021, 322, 114567. [Google Scholar] [CrossRef]
  33. Mohammadi, M.; Benders, S.; Jerschow, A. Nuclear Magnetic Resonance Spin-Lattice Relaxation of Lithium Ions in Aqueous Solution by NMR and Molecular Dynamics. J. Chem. Phys. 2020, 153, 184502. [Google Scholar] [CrossRef] [PubMed]
  34. Hayamizu, K.; Tsuzuki, S.; Seki, S.; Fujii, K.; Suenaga, M.; Umebayashi, Y. Studies on the Translational and Rotational Motions of Ionic Liquids Composed of N-Methyl-N-Propyl-Pyrrolidinium (P13) Cation and Bis(Trifluoromethanesulfonyl)Amide and Bis(Fluorosulfonyl)Amide Anions and Their Binary Systems Including Lithium Salts. J. Chem. Phys. 2010, 133, 194505. [Google Scholar] [CrossRef] [PubMed]
  35. Hayamizu, K.; Tsuzuki, S.; Seki, S.; Umebayashi, Y. Nuclear Magnetic Resonance Studies on the Rotational and Translational Motions of Ionic Liquids Composed of 1-Ethyl-3-Methylimidazolium Cation and Bis(Trifluoromethanesulfonyl)Amide and Bis(Fluorosulfonyl)Amide Anions and Their Binary Systems Including Li. J. Chem. Phys. 2011, 135, 084505. [Google Scholar] [CrossRef] [PubMed]
  36. Bayley, P.M.; Best, A.S.; MacFarlane, D.R.; Forsyth, M. The Effect of Coordinating and Non-Coordinating Additives on the Transport Properties in Ionic Liquid Electrolytes for Lithium Batteries. Phys. Chem. Chem. Phys. 2011, 13, 4632–4640. [Google Scholar] [CrossRef] [PubMed]
  37. Neuhaus, D.; Williamson, M. The Nuclear Overhauser Effect in Structural and Conformational Analysis, 2nd ed.; Wiley: New York, NY, USA, 2000. [Google Scholar]
  38. Hayamizu, K.; Tsuzuki, S.; Seki, S.; Umebayashi, Y. Multinuclear NMR Studies on Translational and Rotational Motion for Two Ionic Liquids Composed of BF4 Anion. J. Phys. Chem. B 2012, 116, 11284–11291. [Google Scholar] [CrossRef]
  39. Alam, T.M.; Dreyer, D.R.; Bielwaski, C.W.; Ruoff, R.S. Measuring Molecular Dynamics and Activation Energies for Quaternary Acyclic Ammonium and Cyclic Pyrrolidinium Ionic Liquids Using 14N NMR Spectroscopy. J. Phys. Chem. A 2011, 115, 4307–4316. [Google Scholar] [CrossRef]
  40. Carof, A.; Salanne, M.; Charpentier, T.; Rotenberg, B. Collective Water Dynamics in the First Solvation Shell Drive the NMR Relaxation of Aqueous Quadrupolar Cations. J. Chem. Phys. 2016, 145, 124508. [Google Scholar] [CrossRef] [Green Version]
  41. Carof, A.; Salanne, M.; Charpentier, T.; Rotenberg, B. On the Microscopic Fluctuations Driving the NMR Relaxation of Quadrupolar Ions in Water. J. Chem. Phys. 2015, 143, 194504. [Google Scholar] [CrossRef]
Figure 1. 1H and 7Li chemical shift variations of ChCl:U-LiCl(0) (top, grey), ChCl:U-LiCl(0.8) (middle, red), and ChCl:U-LiCl(1.0) (bottom, blue), as a function of temperature. Maximum errors are estimated to be 2% for 1H and 7Li, and 4% for OH protons.
Figure 1. 1H and 7Li chemical shift variations of ChCl:U-LiCl(0) (top, grey), ChCl:U-LiCl(0.8) (middle, red), and ChCl:U-LiCl(1.0) (bottom, blue), as a function of temperature. Maximum errors are estimated to be 2% for 1H and 7Li, and 4% for OH protons.
Materials 15 07459 g001
Figure 2. (a) Diffusion coefficients obtained for Ch (triangle), U (square) and Li (star) in the samples ChCl:U-LiCl(0) (grey), ChCl:U-LiCl(0.8) (red), and ChCl:U-LiCl(1.0) (blue), as a function of temperature. The self-diffusion of CH3 is used here as representative for Ch, all cholinium protons exhibiting the same diffusivity. Maximum errors are estimated to be 3% for 1H and 5% for 7Li. (b) Apparent translational activation energy E a t r a n s l obtained from the Arrhenius plots of diffusion data of for Ch (triangle), U (square) and Li (star), as a function of the LiCl content (0 mol/kg = grey, 0.8 mol/kg = red, and 1.0 mol/kg = blue).
Figure 2. (a) Diffusion coefficients obtained for Ch (triangle), U (square) and Li (star) in the samples ChCl:U-LiCl(0) (grey), ChCl:U-LiCl(0.8) (red), and ChCl:U-LiCl(1.0) (blue), as a function of temperature. The self-diffusion of CH3 is used here as representative for Ch, all cholinium protons exhibiting the same diffusivity. Maximum errors are estimated to be 3% for 1H and 5% for 7Li. (b) Apparent translational activation energy E a t r a n s l obtained from the Arrhenius plots of diffusion data of for Ch (triangle), U (square) and Li (star), as a function of the LiCl content (0 mol/kg = grey, 0.8 mol/kg = red, and 1.0 mol/kg = blue).
Materials 15 07459 g002
Figure 3. Apparent transference number for lithium ions in the two Li-doped systems (0.8 mol/kg = red, and 1.0 mol/kg = blue), assuming the self-diffusion coefficient of chloride equal to lithium (filled symbols) or urea (empty symbols), as a function of temperature.
Figure 3. Apparent transference number for lithium ions in the two Li-doped systems (0.8 mol/kg = red, and 1.0 mol/kg = blue), assuming the self-diffusion coefficient of chloride equal to lithium (filled symbols) or urea (empty symbols), as a function of temperature.
Materials 15 07459 g003
Figure 4. 1H and 7Li correlation time τ C calculated as a function of temperature in the samples ChCl:U-LiCl(0) (top, grey), ChCl:U-LiCl(0.8) (middle, red), and ChCl:U-LiCl(1.0) (bottom, blue).
Figure 4. 1H and 7Li correlation time τ C calculated as a function of temperature in the samples ChCl:U-LiCl(0) (top, grey), ChCl:U-LiCl(0.8) (middle, red), and ChCl:U-LiCl(1.0) (bottom, blue).
Materials 15 07459 g004
Figure 5. Average single-jump distance for a lithium ion in the two Li-doped systems (0.8 mol/kg = red, and 1.0 mol/kg = blue), as a function of temperature.
Figure 5. Average single-jump distance for a lithium ion in the two Li-doped systems (0.8 mol/kg = red, and 1.0 mol/kg = blue), as a function of temperature.
Materials 15 07459 g005
Table 1. Samples used in this work.
Table 1. Samples used in this work.
SampleLiCl Content (mol/kg)LiCl Content (wt%)
ChCl:U-LiCl(0)00
ChCl:U-LiCl(0.8)0.83.2
ChCl:U-LiCl(1.0)1.03.9
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

de Araujo Lima e Souza, G.; Di Pietro, M.E.; Castiglione, F.; Vanoli, V.; Mele, A. Insights into the Effect of Lithium Doping on the Deep Eutectic Solvent Choline Chloride:Urea. Materials 2022, 15, 7459. https://doi.org/10.3390/ma15217459

AMA Style

de Araujo Lima e Souza G, Di Pietro ME, Castiglione F, Vanoli V, Mele A. Insights into the Effect of Lithium Doping on the Deep Eutectic Solvent Choline Chloride:Urea. Materials. 2022; 15(21):7459. https://doi.org/10.3390/ma15217459

Chicago/Turabian Style

de Araujo Lima e Souza, Giselle, Maria Enrica Di Pietro, Franca Castiglione, Valeria Vanoli, and Andrea Mele. 2022. "Insights into the Effect of Lithium Doping on the Deep Eutectic Solvent Choline Chloride:Urea" Materials 15, no. 21: 7459. https://doi.org/10.3390/ma15217459

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop