Significance
The development of cost-effective batteries for long-duration grid scale energy storage will be accelerated using frameworks to rapidly screen and select battery components. Herein, we show that the solvent reorganization energy calculated from the Born equation (with reference to an electrolyte’s composition) is predictive of the electrolytes’ device level performance. This descriptor was found to correlate with key transport and kinetic properties over a range of electrolyte compositions and pH values, succinctly capturing the multicomponent interactions between the electrolyte salts and solvent. This enables the initial high-throughput screening of electrolyte candidates with minimal experimentation. Applied to aqueous redox flow batteries employing organic redox active species, we predict high-performance electrolyte compositions, enabling significantly enhanced device performance.
Keywords: redox flow battery, Marcus–Hush theory, Thiele modulus, effectiveness factor, organic active species
Abstract
Organic and organometallic reactants in aqueous electrolytes, being composed of earth-abundant elements, are promising redox active candidates for cost-effective organic redox flow batteries (ORFBs). Various compounds of ferrocene and methyl viologen have been examined as promising redox actives for this application. Herein, we examined the influence of the electrolyte pH and the salt anion on model redox active organic cations, bis((3-trimethylammonio) propyl)- ferrocene dichloride (BTMAP-Fc) and bis(3-trimethylammonio) propyl viologen tetrachloride (BTMAP-Vi), which have exhibited excellent cycling stability and capacity retention at ≥1.00 M concentration [E. S. Beh, et al. ACS Energy Lett. 2, 639–644 (2017)]. We examined the solvation shell around BTMAP-Fc and BTMAP-Vi at acidic and neutral pH with SO42-, Cl−, and CH3SO3− counterions and elucidated their impact on cation diffusion coefficient, first electron transfer rate constant, and thereby the electrochemical Thiele modulus. The electrochemical Thiele modulus was found to be exponentially correlated with the solvent reorganizational energy (λ) in both neutral and acidic pH. Thus, λ is proposed as a universal descriptor and selection criteria for organic redox flow battery electrolyte compositions. In the specific case of the BTMAP-Fc/BTMAP-Vi ORFB, low pH electrolytes with methanesulfonate or chloride counterions were identified as offering the best balance of transport and kinetic requirements.
The market-driven adoption of renewable power sources such as solar and wind due to the decrease in the unsubsidized levelized cost of electricity (currently cheaper than every source except natural gas) necessitates the adoption of grid scale energy storage to ensure grid reliability and resiliency (1). The stringent requirements for the economic viability of these sources, such as a cost of <$100/KWh (2), levelized cost of storage of <$0.05/kWh-cycle (3), and 10 to 100 h of discharge (4) at rated power, imposes significant constraints on the possible technologies that meet these targets. Redox flow batteries (RFBs), which decouple the power and energy output of a battery by storing the redox active material outside the battery itself, are a promising grid scale energy storage technology (5). Various RFBs employing elemental, metallic actives such as Fe-Cr (6), all-V (7, 8), Zn-Ce (9), and V-Ce (10, 11) have been the subject of active research over the past half century. The capacity of these systems is determined by the actives’ solubility (typically ∼1 M) and the number of electrons transferred per redox active ion. The typical 1 M concentration and the single electron transfer possible with a vast majority of these systems caps their theoretical specific capacity at 26.8 Ah ⋅ L−1 with the energy density being restricted to a maximum of ∼35 Wh ⋅ L−1 (assuming a generous ∼1.5 V operation). The cost-conscious decision to utilize the cheapest solvent, water, focuses developmental efforts on improving the solubility and number of electrons transferred within the aqueous electrochemical window.
Recent developments in utilizing water-soluble organic/organometallic actives, some with multiple redox centers, has led to the demonstration of organic RFBs (ORFBs) with reasonably long cycle life and all-V RFB-like capacity. The key incentive for moving to these organic redox active species is the potential for multielectron redox processes at each electrode and the >1 M solubility limit of these organic species. For example, a two-electron transfer redox process at each electrode and a 1.5 M solubility limit for a hypothetical organic active species yields a theoretical specific capacity of 80.4 Ah ⋅ L−1. To realize this potential, a variety of RFBs based on methyl viologen (12–16), (2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (12–14), ferrocene (15, 16), various quinones (17–20), ferricyanide (17), and alloxazine (21) actives in acidic, neutral, and alkaline pH conditions have been developed. The performance of some of these RFBs have resulted in plans for large-scale deployment of such systems. In other systems, extraordinarily long cycles lives in the order of centuries (200 y to 50% capacity loss) have been projected (22). These organic actives are invariably composed of chemicals that are low cost at scale (e.g., the viologen paraquat is an extensively used herbicide). Thus, after initial capital investments in scaled-up production, the cost of these materials would be lower than typical inorganic RFB actives (23). This is due to the fact that these organic actives can result in RFBs with higher specific energy, thereby lowering the cost on a capacity normalized basis (24).
Given the stringent cost consideration for grid scale deployment, any ORFB needs to maximize its energy density and power density. Engineering solutions exist to improve most factors that influence energy density and power density, like the reactant concentration, current density, and cell/stack component electrical resistance, but a key parameter, the open circuit voltage (OCV), is inherently a function of the redox potentials of the solvated actives. A selected chemistry with low discharge polarization and high OCV would exhibit superior energy and power densities. Thus, we examined the effect of the supporting electrolyte components on the transport and kinetic properties of these organic actives, as it is the solvated ions that take part in the interactions of interest.
We selected a configuration with bis(3-trimethylammonio) propyl)ferrocene dichloride (BTMAP-Fc) catholyte coupled with bis(3-trimethylammonio)propyl viologen tetrachloride (BTMAP-Vi) anolyte in an aqueous ORFB as our model system. At neutral pH, this system has shown high capacity, high stability, and low crossover (16). The present study examined the effect of pH and counter ion selection on the transport and kinetics of the model organic actives BTMAP-Fc and BTMAP-Vi. Given concerns with alkaline stability, only acidic (H2SO4, HCl, and CH3SO3H) and neutral (Na2SO4, NaCl, and NaCH3SO3) pH aqueous electrolytes were considered (25). The schematic of the ORFB and the various electrolyte compositions considered are depicted in Fig. 1. The effect of the pH and counter anion on the diffusion coefficient, hydrodynamic radius, first electron transfer rate constant, and solvent reorganization energy () was examined using standard electrochemical methods. The electrochemical Thiele modulus (), combining the diffusion coefficient and first electron transfer rate constant, was found to be strongly correlated with the solvent reorganization energy (). Detailed discussions of in the context of Marcus–Hush (M-H) kinetics and its advantages over the widely used Butler–Volmer (B-V) kinetics formulation have been presented in our prior publications (26–29). We have previously shown that similar strong predictive correlations exist between and in nonaqueous electrolytes (29). Having established that such correlations hold for aqueous and nonaqueous electrolytes at up to 1 M dissolved salt concentrations, extensions of this framework to highly concentrated water-in-salt, ionic liquid, and deep eutectic solvent-based electrolytes could be fruitful. The optimal range of electrolyte that resulted in high OCVs was determined, and recommendations on both optimal electrolyte pH and choice of counterion are provided. Although Marcus theory predicts an inverse relationship between the heterogeneous electron transfer kinetics and the solvent reorganization energy, here, we conclusively demonstrate the practical predictive applicability of M-H theory to electrochemical device engineering by showing how Marcus theory parameters are related to (and, hence, predictive of) device level metrics like the Thiele modulus.
Fig. 1.
Schematic of an ORFB assembled with BTMAP-Fc catholyte and BTMAP-Vi anolyte incorporating an anion exchange membrane separator. The pH and anion combinations in the supporting electrolyte are tabulated below the schematic. The electron and ion flow during charging is depicted using a solid line, and the equivalent flow during the discharge process is depicted using dashed lines.
Materials and Methods
Chemicals.
The electrochemical measurements reported herein were carried out in 1 M aqueous solutions of BTMAP-Fc and BTMAP-Vi with six different supporting electrolytes. The supporting electrolytes used were sulfuric acid (H2SO4) (Sigma-Aldrich, ≥99%), HCl (Sigma-Aldrich, American Chemical Society [ACS] reagent 37%), CH3SO3H (Sigma-Aldrich, ≥99%), Na2SO4 (Sigma-Aldrich, ACS reagent ≥99%), NaCl (Sigma-Aldrich, ACS reagent ≥99%), and NaCH3SO3 (Sigma-Aldrich, ≥98%). All the supporting electrolytes used in the study were at 1 M concentration.
Synthesis of Model Organic Redox Active Species.
The organic redox active species BTMAP-Fc and BTAP-Vi, detailed in the work of Beh et al. (16), were chosen as the model compounds (their respective structures are depicted in Fig. 1). The compounds were synthesized as detailed in the supplementary information of that paper, and successful synthesis was confirmed by H1 NMR. All air and moisture sensitive steps in the synthesis were carried out in a MBraun Ar-filled glove box with <0.5 ppm of O2 and H2O. The salts were prepared and stored in their Cl− forms and suitably ion exchanged using amberlite ion exchange resins into the desired form before electrochemical measurements. An additional supporting electrolyte was added following ion exchange.
Viscosity Measurements.
The viscosity of all the six electrolytes was measured using Ubbelohde viscometer. The concentration of all the solutions were 1 M. All the measurements were done at 25 °C. The viscometer was vertically inserted into a constant temperature bath and charged by pouring sample in lower meniscus. Suction was applied to the tube, and efflux time in the capillary was noted. Kinematic viscosity was calculated by multiplying efflux time to viscometer constant. From kinematic viscosity, dynamic viscosity was calculated by dividing the kinematic viscosity by density of the solution. The viscosities are listed in SI Appendix, Table S3.
Electrochemical Measurements.
The synthesized BTMAP-Fc-Cl and BTMAP-Vi-Cl salts were stored in a MBraun argon-filled glove box with H2O and O2 levels <0.5 ppm to prevent moisture absorption. The salts were weighed and removed from the glove box prior to each experimental run, and the electrolytes were prepared in a fume hood.
The electrochemical measurements were carried out in a setup that consisted of a small-volume electrochemical cell (Pine Instruments, RRPG223) with polytetrafluoroethylene (PTFE) stoppers and openings for working, counter, and reference electrodes and gas purge inlet and exit. A 3-mm-diameter glassy carbon (GC) disk was used as the working electrode; the counter electrode consisted of a Pt mesh attached to a Pt wire, and the reference electrode was Ag/AgCl. The electrochemical measurements were performed using a Gamry Instruments Series-G single channel potentiostat. The GC was chosen as a working electrode because it is made up of a noncatalytic material in the potential window of interest and will equally reduce or promote surface adsorption of electroactive species in all the six electrolytes such that kinetic parameters can be compared.
The electrolyte’s conductivity measurements were carried out with a two-electrode cell using electrochemical impedance spectroscopy with a Solartron analytical potentiostat (1470 Cell test system) connected to a 1252A Frequency Response Analyzer. A total of 30 mL of solution was placed in a beaker with two titanium fiber felt electrodes (Fuel Cell Store) separated by a distance of 1.6 cm with an immersed area of 1.2 cm2. The potentiostat was used to measure the impedance response in the frequency range 300,000 to 1,000 Hz. The high-frequency resistance was estimated from Bode plots corresponding to a phase angle of zero (SI Appendix, Fig. S5). Conductivity was calculated using the following equation:
| [1] |
where κ is conductivity in Siemens per centimeter (S/cm), l is distance between electrodes, A is cross-sectional area of sample, and R is resistance.
Results and Discussion
Electrochemical Characterization.
The redox characteristics of the BTMAP-Fc and BTMAP-Vi actives were initially evaluated using cyclic voltammetry (CV) as depicted in Fig. 2. The CV characteristics (after iR correction), such as the cathodic and anodic peak potentials (Epc and Epa, respectively), peak separation (), half-wave potential (), and the transfer coefficient (), are tabulated for both the catholyte and anolyte compositions in Tables 1 and 2, respectively. The transfer coefficient, , was calculated using the following equation (30):
| [2] |
where R (universal gas constant) = 8.314 J ⋅ mol-1 ⋅ K−1, T is the temperature in K, F (Faraday’s constant) = 96,485 C ⋅ mol−1 , and is the half-peak potential. The pH, ionic strength, and counterions were found to significantly influence the half-wave potentials (calculated as ), which were taken to closely approximate the formal potentials of the redox species (10, 30). is approximated as the formal potential following Bard and coworkers (30) because at a low scan rate, the peak current ratio is approximately equal to 1 (SI Appendix, Table S6). Low pH electrolytes exhibited a greater difference between the half-wave potentials of BTMAP-Fc and BTMAP-Vi (Table 3), suggesting that ORFBs employing low pH electrolytes would exhibit higher OCV and, hence, greater peak power densities (all other overpotential contributions being equal). The magnitudes of the anodic and cathodic currents were invariably found to be higher in low-pH electrolytes, except in the case of BTMAP-Fc with methanesulfonate counterion, where the pH = 7 electrolyte exhibited higher currents. All redox species were found to exhibit >59 mV overpotential for both anodic and cathodic reactions, indicating that the reactions are not fully reversible (30). This indicated significant charge and discharge overpotentials. Thus, any increases in OCV conferred by the use of low-pH electrolytes were deemed to be advantageous because, all overpotentials being equal, the higher OCV cell will exhibit higher operating voltage. Furthermore, low-pH electrolytes were invariably found to exhibit higher conductivity compared to pH = 7 electrolytes (SI Appendix, Table S1). H2SO4 was found to exhibit the highest conductivity. Since the other overpotential contributions from ohmic losses and concentration polarization would be the same between systems (given the same separator and pumping rate), the choice of low-pH electrolytes is advantageous from the perspective of lowering ohmic losses.
Fig. 2.
CVs of BTMAP-Fc and BTMAP-Vi aqueous solutions with the following supporting electrolytes: H2SO4 (A), Na2SO4 (B), HCl (C), NaCl (D), CH3SO3H (E), and NaCH3SO3 (F). All CVs were measured at 225 mV ⋅ s−1 and represent the average of three scans. WE: GC; CE: Pt; RE: Ag/AgCl.
Table 1.
Voltammetric properties for BTMAP-Fc containing electrolytes obtained from CVs in Fig. 2
| Electrolyte | Epc (V) | Epa (V) | ΔEp (V) | E1/2 (V) | α |
| H2SO4 | 0.43 | 0.38 | 0.06 | 0.41 | 0.71 |
| HCl | 0.45 | 0.37 | 0.08 | 0.41 | 0.74 |
| CH3SO3H | 0.48 | 0.40 | 0.08 | 0.44 | 0.67 |
| Na2SO4 | 0.49 | 0.41 | 0.08 | 0.45 | 0.77 |
| NaCl | 0.47 | 0.40 | 0.08 | 0.44 | 0.67 |
| NaCH3SO3 | 0.51 | 0.41 | 0.10 | 0.46 | 0.38 |
Table 2.
Voltammetric properties for BTMAP-Vi containing electrolytes obtained from CVs in Fig. 2
| Electrolyte | Epc (V) | Epa (V) | ΔEp (V) | E1/2 (V) | α |
| H2SO4 | −0.46 | −0.35 | −0.11 | −0.41 | 0.53 |
| HCl | −0.43 | −0.35 | −0.07 | −0.39 | 0.59 |
| CH3SO3H | −0.43 | −0.32 | −0.11 | −0.38 | 0.54 |
| Na2SO4 | −0.43 | −0.28 | −0.15 | −0.36 | 0.31 |
| NaCl | −0.37 | −0.30 | −0.07 | −0.33 | 0.52 |
| NaCH3SO3 | −0.39 | −0.27 | −0.12 | −0.33 | 0.38 |
Table 3.
Supporting electrolyte-dependent OCV values for BTMAP-Fc/BTMAP-Vi ORFBs
| Electrolyte (1 M) | Ionic strength | E1/2 (V) (BTMAP-Fc [1 M]) | E1/2 (V) (BTMAP-Vi [1 M]) | OCV (V) |
| H2SO4 | 3.00 | 0.41 | −0.41 | 0.82 ± 0.01 |
| HCl | 1.00 | 0.41 | −0.39 | 0.80 ± 0.02 |
| CH3SO3H | 1.00 | 0.44 | −0.38 | 0.82 ± 0.01 |
| Na2SO4 | 3.00 | 0.45 | −0.36 | 0.80 ± 0.01 |
| NaCl | 1.00 | 0.44 | −0.33 | 0.77 ± 0.00 |
| NaCH3SO3 | 1.00 | 0.46 | −0.33 | 0.79 ± 0.01 |
The transfer coefficient, , indicates the tendency of a given redox species in its transition state to go to its oxidized or reduced form for the one-electron transfer reactions under consideration here (30). A value of = 0.5 indicates that oxidation and reduction (i.e., the anodic and cathodic reactions) were equally facile. The values listed in Table 1 are for the oxidation (anodic process) of BTMAP-Fc, while values in Table 2 are for the reduction (cathodic process) of BTMAP-Vi since the organic actives were synthesized in their “discharged” state. Tafel analysis using these values provide an indication of the overpotentials required to drive the anodic and cathodic reactions. The Tafel equation is as follows:
| [3] |
where
| [4] |
| [5] |
Here, i0 is the exchange current density in milliampere per square centimetre (mA cm−2). The value of the Tafel slope for a one-electron transfer reaction with the is 118 mV ⋅ dec−1. In this ideal case, the ratio of anodic and cathodic Tafel slopes is 1 and their difference is 0. Since the overall reactions for both the catholyte and anolyte are a one-step and one-electron transfer, the sum of cathodic and anodic transfer coefficients (31). A higher Tafel slope value indicates that greater overpotential is required to drive a given increase in current (and hence rate of reaction). The absolute values of the anodic and cathodic Tafel slopes for all compositions of the anolyte and catholyte are tabulated in Table 4. The NaCl supported electrolytes came closest to the ideal Tafel slope ratio of one and exhibited the lowest average difference in Tafel slopes for both the anolyte and catholyte. Another promising composition (based on the same analysis) was the CH3SO3H-supported electrolyte. Comparing the Tafel slopes with the OCVs in Table 3, high OCV electrolytes with low Tafel slopes (hence, low overpotential) will result in high operating potentials. Thus, we predict that low-pH electrolytes with CH3SO3− counterions will result in the lowest average charge and discharge overpotentials and the highest ORFB OCVs, thereby resulting in the highest peak power values. Having established this experimental baseline, the fundamental cation transport and heterogenous electron kinetics were examined to determine possible correlation with the solvent reorganization energy. Similar to our previous studies in nonaqueous systems, if can serve as a universal nonheuristic descriptor for selecting high-performance electrolyte compositions, this would significantly reduce the amount of trial-and-error experimentation with various electrolyte compositions required to arrive at the optimum.
Table 4.
Calculated anodic and cathodic Tafel slopes for BTMAP-Fc and BTMAP-Vi electrolytes
| Electrolyte | BTMAP-Fc | BTMAP-Vi | ||
| Anodic | Cathodic | Anodic | Cathodic | |
| (mV ⋅ dec−1) | (mV ⋅ dec−1) | (mV ⋅ dec−1) | (mV ⋅ dec−1) | |
| H2SO4 | 80.58 ± 1.09 | 222.35 ± 7.93 | 126.65 ± 1.98 | 110.80 ± 1.48 |
| HCl | 80.20 ± 0.79 | 224.77 ± 6.04 | 144.28 ± 3.32 | 100.18 ± 1.60 |
| CH3SO3H | 87.74 ± 0.93 | 181.06 ± 4.05 | 126.61 ± 1.46 | 110.77 ± 1.14 |
| Na2SO4 | 75.10 ± 0.69 | 277.77 ± 9.19 | 87.74 ± 0.93 | 181.06 ± 4.05 |
| NaCl | 88.16 ± 0.62 | 179.08 ± 2.56 | 119.76 ± 1.34 | 116.61 ± 1.24 |
| NaCH3SO3 | 162.92 ± 4.51 | 92.83 ± 1.41 | 93.76 ± 0.70 | 159.70 ± 2.04 |
Diffusion Coefficient and Hydrodynamic Radii of the Solvated Model Organic Cations.
CVs at different potential scan rates were obtained as shown in SI Appendix, Figs. S1 and S2. These voltammograms displayed the characteristic 30/α mV shift in the reduction peak position of the CV for a 10-fold increase in scan rate. Based on the peak separations tabulated in Tables 1 and 2, the Nicholson–Shain (N-S) equation (32, 33) was used to relate the peak currents to the scan rate. The relationship is as follows:
| [6] |
where ip (A) is the peak current, n is the number of electrons involved in the electrode reaction (n = 1), α is the charge transfer coefficient, A (squared centimeters) is the electrode area, C in mole per cubic centimeters (mol ⋅ cm−3) is the concentration of electroactive organic compounds, T (K) is the temperature, D (squared centimeters per second) is the diffusion coefficient of electroactive organic compounds, and (V ⋅ s−1) is the scan rate at which the CV was recorded. The N-S plots are depicted in SI Appendix, Figs. S3 and S4. The plots were linear with square root of scan rate as expected, and their slopes were used to calculate the diffusion coefficient of electroactive organic compound. The redox reactions considered in this work do not involve any bond formation or breakage. Under such conditions, the variation in the radius of the reduced and oxidized species can be quite small (30) and results in (which is the case at small scan rates for the system under consideration). Thus, the diffusion coefficient of the oxidized and reduced state is taken to be the same [following Bond and coworkers (34)], and and . The diffusion coefficients of the catholyte and anolyte in the various supporting electrolytes considered are depicted in Figs. 3A and 4A, respectively. The hydrodynamic radius of the solvated ions in solution is directly correlated with the mean free path and typically exhibits an inverse relationship with diffusion coefficient as per the Stokes–Einsten relation (35). According to the Stokes–Einstein equation,
| [7] |
where D (cm2 ⋅ s−1) is diffusion coefficient, kB (1.38064852 × 10−23 m2 ⋅ kg ⋅ s−2 ⋅ K−1) is Boltzmann’s constant, T (298 K) is temperature, η (Pa s) is the dynamic viscosity, and rhyd (meters) is the hydrodynamic radius.
Fig. 3.
Transport and kinetic properties of BTMAP-Fc aqueous solution in acidic electrolytes (H2SO4, HCl, and CH3SO3H) and neutral electrolytes (Na2SO4, NaCl, and NaCH3SO3). (A) Diffusion coefficient, (B) hydrodynamic radius, (C) rate constant, and (D) solvation reorganization overpotential. The error bars represent SE from at least three independent measurements. The corresponding numerical values are tabulated in SI Appendix, Table S3.
Fig. 4.
Transport and kinetic properties of BTMAP-Vi aqueous solution in acidic electrolytes (H2SO4, HCl, and CH3SO3H) and neutral electrolytes (Na2SO4, NaCl, and NaCH3SO3). (A) Diffusion coefficient, (B) hydrodynamic radius, (C) rate constant, and (D) solvation reorganization overpotential. The error bars represent SE from at least three independent measurements. The corresponding numerical values are tabulated in SI Appendix, Table S3.
The diffusion coefficient is also influenced by electrolyte composition–dependent intermolecular forces, which are macroscopically measured as the electrolyte’s viscosity. The viscosities of all the six electrolytes were measured using an Ubbelohde viscometer, listed in SI Appendix, Table S2. The viscosity increased with change in counterion in the order SO42− > CH3SO3− > Cl−. The hydrodynamic radius in different supporting electrolytes is depicted in Figs. 3B and 4B. The expectation that the hydrodynamic radius would increase in the order Cl− < HSO4− < CH3SO3− with increasing counterion radii was not borne out for either cation. The diffusion coefficient was found to exhibit an inverse trend with the hydrodynamic radii for both cations as expected from the Stokes–Einstein equation. Nevertheless, the ratios of cation D between electrolytes were not the same as ratios of rhyd, indicating the influence of the viscosity. Also, while the D in acidic and neutral media were of the same order of magnitude for BTMAP-Fc, D decreased by an order of magnitude for BTMAP-Vi when transitioning from acidic to neutral media. The decrease in D resulted in a concomitant increase in rhyd. This is attributed to greater solvation by polar water molecules of the BTMAP-Vi (with four N+ charge centers) in neutral media. We attribute the lack of a similar effect in case of BTMAP-Fc to the + charges from the two N+ charge centers being restricted to one end of the cation as opposed to the more uniform distribution in BTMAP-Vi. Given this complex interplay of factors, the organic cation diffusion coefficient cannot be readily increased by just choosing counterions of smaller ionic radii. There is therefore a need to experimentally determine D for new organic active–based electrolyte compositions at present.
Electrochemical Kinetics and the Solvation Overpotential.
The rate constant was calculated using the Kochi and Klingler method (36). A transient method (CV) was employed to study this system given fears of possible side reactions. BTMAP-Vi electrolyte undergoes dimerization leading to the formation of a charged transfer complex (37–39), and the ferrocene or ferrocenium may adsorb on the electrode and poison it (40, 41). We have previously screened electrolytes by applying M-H theory to steady-state voltammetry data when such side-reaction considerations were not warranted (29). Kochi and Klingler evaluated the standard rate constant [k (cm ⋅ s−1)] from CV curves by means of the relationship:
| [8] |
where α is the transfer coefficient for the redox reaction (listed in Tables 1 and 2), and ΔEp (V) is the peak-to-peak separation of the anodic and cathodic waves. CV measurements on both reactants gave an oxidation rate constant of order 10−1 cm ⋅ s−1 for BTMAP-Fc and a reduction rate constant of order 10−2 cm ⋅ s−1 for BTMAP-Vi in various supporting electrolytes. Thus, BTMAP-Vi was identified as the rate-limiting side. These reactions are much faster than those of common inorganic species and are also faster than those of most other organic or organometallic reactants that have been used in RFBs, as tabulated in SI Appendix, Table S4. These measurements utilized a noncatalytic GC electrode to obtain a baseline rate constant value. Based on the rate-limiting electrode reaction, recommendations can then be made for the use of catalysts. We deliberately avoided the use of a catalytic electrode, as we have previously found that catalytic electrodes may actually be detrimental to outer-sphere electron transfer processes and lead to incorrect identification of the rate-limiting electrode reaction (28). Given the high-rate constant values for BTMAP-Fc and BTMAP-Vi redox reactions compared to typical RFB electrodes, the use of catalysts may not be warranted at either electrode. BTMAP-Vi being rate limiting suggests (at first glance) that the HCl-supporting electrolyte, which results in the highest rate constant for BTMAP-Vi, should be chosen for the overall system. However, the overall combination of bulky cations and fast first electron transfer kinetics results in BTMAP-FC/BTMAP-Vi ORFBs being mass transport limited under some conditions. Thus, a judicious optimum of D and k is needed. Given the counterintuitive trends in D and fast first electron transfer to the solvated cations, we examined the solvation energy for further insights.
The solvent reorganization energy (), which is the amount of energy required to change the solvation shell of the reactants to that of the product during an electrochemical reaction, plays a prominent part in the M-H kinetic formulation (42). There are two contributions to λ, one from the changes in the intramolecular vibrations (the inner-sphere reorganization energy, λi) and the second from the changes in solvent orientation coordinates (the outer sphere reorganization energy, lo) (30, 43–45):
| [9] |
The inner component is described by summing over the normal vibrational modes of the species using molecular theory:
| [10] |
where kj is the force constant for each oscillator, and q is the displacement in the normal mode coordinates.
The outer sphere reorganization energy is usually approximated by modeling the reactants and products as spheres and the solvent as a dielectric continuum (Born theory), as seen in the work of Bard and Faulkner (30), Compton et al. (43, 44, 46), Savéant and Costentin (47), Marcus (43, 46, 48), and Chidsey (43, 48). The Born equation (49) can be used for estimating the electrostatic component of Gibbs free energy of solvation of an ion and was used to estimate . It is an electrostatic model that treats the solvent as a continuous dielectric medium and is as follows:
| [11] |
where e is the elementary charge, εo is the permittivity of free space, ao is the radius of the reactant, εop is optical permittivity equal to the square of refractive index of solvent, εr is relative permittivity, and d is the distance from the reactant surface to the metal surface.
As, d > ao,
| [12] |
When the oxidizing and reducing species are similar in structure, internal reorganizational energy is smaller than λo. Therefore, we can use λo as an estimate for reorganizational energy.
| [13] |
where NA (6.02214076 × 1023 mol−1) is Avogadro’s constant, z is charge of the ion, e (1.6022 × 10−19 C) is elementary charge, ε0 (8.8541878128 × 10−12 F ⋅ m−1) is permittivity of free space, rhyd is effective radius of ion (i.e., hydrodynamic radius), εop (1.775) is optical permittivity, and εr (78.4) is relative permittivity of the solvent (i.e., water). λ0 encompasses both the ease of solvation of a given species and the ease of oxidizing or reducing it by calculating |λo/F|, the solvation reorganization overpotential required to rearrange or break the solvation shell of a given reactant. The |λo/F| values for BTMAP-Fc and BTMAP-Vi are provided in Figs. 3D and4D, respectively. Assuming all other overpotential contributions are equal for a given cation in various supporting electrolytes, |λo/F| will then determine the overall overpotential and, hence, the voltage efficiency of a full ORFB cell. The effective radii of the cationic complexes were understood to be a function of both the radii of the counterions and also their charge density. Thus, a nominally larger counterion with a high charge density results in an overall smaller complex by reducing the distance between the charge centers. |λ0/F| can also be interpreted to indicate the occurrence of an inner-sphere versus outer-sphere electron transfer process as we have detailed elsewhere (28). Given the large |λo/F| values (greater than even the ) for all electrolyte compositions, outer-sphere electron transfer was indicated in all cases. Given the complex interplay between D, k, and |λo/F|, we elected to simultaneously optimize for both D and k by examining the impact of |λo/F| on the electrochemical Thiele modulus (29).
Electrochemical Thiele Modulus and Ion Solvation.
The transport and kinetic parameters of the various electrolytes considered (depicted in Figs. 3 and 4) were considered in totality by using the electrochemical Thiele modulus (ht) with hydrodynamic radius as the characteristic length (l) (29). The derivation of ht can be found in Sankarasubramanian et al. (29) The value of k (as calculated using the Kochi–Klingler equation) was not mass transport limited and hence satisfied the key boundary condition [from the derivation of the ht relation in our previous work (29)] that the reactant concentration at the surface is finite. The electrochemical Thiele modulus involves a potential dependent k (unlike the constant k in the typical isothermal reaction engineering case) and is related to as follows:
| [14] |
where (V) is the applied overpotential and Cibulk in mole per cubic meters (mol ⋅ m−3) is the concentration of the electroactive species in the electrolyte bulk. The transport dependencies and the overpotential drive force are captured in the groupings A and B which are as follows:
| [15] |
| [16] |
The A and B values are listed in SI Appendix, Table S5. A increased with cation size as expected based on the correlation between D and rhyd. B also increased with cation size as expected based on the correlation between |λo/F| and rhyd. This served as validation of the applicability of this model to the present case. Plots of ln(hT2) versus |λo/F| are depicted in Fig. 5 and exhibited a linear relationship (within experimental error), as expected. To utilize the Thiele modulus to comment on the relative importance of transport versus kinetics in a given system, it is useful to consider the effectiveness factor (; not to be confused with the overpotential, ), which is defined as the ratio of the rate of a reaction when reactant availability is diffusion limited to the rate of reaction at maximum (bulk electrolyte) concentration. The one-dimensional model (29) applied herein allows us to calculate using the relationship for a slab geometry and, hence, (50). A limiting case when ht is small is that ηeff → 1. All the Thiele moduli in our system are in the order of 10−2 (SI Appendix, Table S3), which would result in ηeff → 1, indicating that the redox reactions of both BTMAP-Fc and BTMAP-Vi are occurring at close to their maximum possible reaction rates (i.e., at rates we would expect if the near electrode reactant concentration was close to the bulk concentration). Thus, we predict that increases in the electrolyte flowrates will have minimal influence on the overall power output from the ORFB. The |λo/F| values were found to numerically correlate with the OCV values listed in Table 3. Based on this correlation and Eq. 14, we recommend that (in the absence of experimental data) new electrolyte compositions be chosen based on higher absolute solvation energy calculated from the Born equation (Eq. 13). This will result in higher hT, which, in turn, will result in higher currents at the cell level. For the vast majority of electrochemical reactions, given that the rhyd will be in the order of 10−9 m, hT << 1, resulting in .
Fig. 5.
Correlation between solvated reorganization overpotential (|λο/F|) and the electrochemical Thiele Modulus (incorporating the diffusion coefficient, electron rate constant, and hydrodynamic radius) for (A) BTMAP-Fc and (B) BTMAP-Vi with various counterions in acidic and neutral electrolytes.
Conclusions
The choice of supporting electrolyte greatly influences the transport and kinetic properties in the anolyte and catholyte of an RFB. In this work, we develop a nonheuristic method to select aqueous supporting electrolyte compositions (pH and counterion) for organic redox actives using the solvent reorganization energy () from M-H kinetic theory as a universal descriptor. Overall, solvent reorganizational energy and hence the electrolyte contribution to the overpotential (i.e., solvation overpotential |λo/F|) are found to be inversely correlated with the natural logarithm of the square of the electrochemical Thiele modulus [)]. Based on our analysis, we conclude the following:
-
1)
Effectiveness factor for both the BTMAP-Fc and BTMAP-Vi redox couples indicates that electrolyte flowrate or optimization for better mass transport would result in minimal ORFB performance improvements.
-
2)
Given the exceptionally high first electron transfer rate constants, even on GC, the introduction of catalysts at the ORFB electrodes is not necessary.
-
3)
Both redox couples exhibit outer-sphere electron transfer mechanisms.
-
4)
Barring possible side reactions, we recommend the use of low-pH or CH3SO3H- or HCl-supported electrolytes to improve the energy efficiency of the BTMAP-Fc/BTMAP-Vi ORFB.
The key finding is that the solvation energies calculated from the Born equation fit into our electrochemical Thiele modulus framework. This has the great advantage of allowing for the initial high-throughput screening of electrolyte candidates with minimal experimentation.
Supplementary Material
Acknowledgments
We gratefully acknowledge financial support from the Advanced Research Projects Agency–Energy, the US Department of Energy, under Award No. DE-AR0000768 as part of the Integration and Optimization of Novel Ion Conducting Solids program. We thank Washington University in St. Louis McKelvey School of Engineering and the Roma B. and Raymond H. Wittcoff Distinguished University Professorship for enabling this study. We also thank the High-Resolution NMR Facility, Department of Chemistry, at Washington University in St. Louis for access to the H1-NMR.
Footnotes
The authors declare no competing interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2105889118/-/DCSupplemental.
Data Availability
All study data are included in the article and/or SI Appendix.
References
- 1.US Energy Information Administration , Levelized Cost and Levelized Avoided Cost of New Generation Resources in the Annual Energy Outlook 2020 (US Energy Information Administration, Washington, DC, 2020). [Google Scholar]
- 2.Department of Energy, Funding Opportunity Announcement Advanced Research Projects Agency–Energy (Arpa-E) Integration and Optimization of Novel Ion Conducting Solids (IONICS), Funding Opportunity No. DE-FOA-0001478 (2016).
- 3.Department of Energy, Funding Opportunity Announcement Advanced Research Projects Agency–Energy (Arpa-E) Duration Addition to Electricity Storage (DAYS), Funding Opportunity No. DE-FOA-0001906 (2017).
- 4.Armstrong C. G., Toghill K. E., Stability of molecular radicals in organic non-aqueous redox flow batteries: A mini review. Electrochem. Commun. 91, 19–24 (2018). [Google Scholar]
- 5.Soloveichik G. L., Flow batteries: Current status and trends. Chem. Rev. 115, 11533–11558 (2015). [DOI] [PubMed] [Google Scholar]
- 6.Gahn R.F., Hagedorn N.H., Johnson J.A., “Cycling performance of the iron-chromium redox energy storage system” in Proceedings of the Intersociety Energy Conversion Engineering Conference (Society of Automotive Engineers, Warrendale, PA, 1985), pp. 2.91–2.97. [Google Scholar]
- 7.Rychcik M., Skyllas-Kazacos M., Characteristics of a new all-vanadium redox flow battery. J. Power Sources 22, 59–67 (1988). [Google Scholar]
- 8.Ulaganathan M., et al., Recent advancements in all-vanadium redox flow batteries. Adv. Mater. Interfaces 3, 1–22 (2016). [Google Scholar]
- 9.Leung P. K., Ponce-De-León C., Low C. T. J. J., Shah A. A., Walsh F. C., Characterization of a zinc-cerium flow battery. J. Power Sources 196, 5174–5185 (2011). [Google Scholar]
- 10.Sankarasubramanian S., Zhang Y., Ramani V., Methanesulfonic acid-based electrode-decoupled vanadium-cerium redox flow battery exhibits significantly improved capacity and cycle life. Sustain. Energy Fuels 3, 2417–2425 (2019). [Google Scholar]
- 11.Yun S., Parrondo J., Ramani V., A vanadium-cerium redox flow battery with an anion-exchange membrane separator. ChemPlusChem 80, 412–421 (2015). [Google Scholar]
- 12.Liu T., Wei X., Nie Z., Sprenkle V., Wang W., A total organic aqueous redox flow battery employing a low cost and sustainable methyl viologen anolyte and 4-HO-TEMPO catholyte. Adv. Energy Mater. 6, 1501449 (2016). [Google Scholar]
- 13.Janoschka T., et al., An aqueous, polymer-based redox-flow battery using non-corrosive, safe, and low-cost materials. Nature 527, 78–81 (2015). [DOI] [PubMed] [Google Scholar]
- 14.Janoschka T., Martin N., Hager M. D., Schubert U. S., An aqueous redox-flow battery with high capacity and power: The TEMPTMA/MV system. Angew. Chem. Int. Ed. Engl. 55, 14427–14430 (2016). [DOI] [PubMed] [Google Scholar]
- 15.Hu B., DeBruler C., Rhodes Z., Liu T. L., Long-cycling aqueous organic redox flow battery (AORFB) toward sustainable and safe energy storage. J. Am. Chem. Soc. 139, 1207–1214 (2017). [DOI] [PubMed] [Google Scholar]
- 16.Beh E. S., et al., A neutral PH aqueous organic- organometallic redox flow battery with extremely high capacity retention. ACS Energy Lett. 2, 639–644 (2017). [Google Scholar]
- 17.Lin K., et al., Alkaline quinone flow battery. Science 349, 1529–1532 (2015). [DOI] [PubMed] [Google Scholar]
- 18.Huskinson B., et al., A metal-free organic-inorganic aqueous flow battery. Nature 505, 195–198 (2014). [DOI] [PubMed] [Google Scholar]
- 19.Yang B., et al., High-performance aqueous organic flow battery with quinone-based redox couples at both electrodes. J. Electrochem. Soc. 163, A1442–A1449 (2016). [Google Scholar]
- 20.Yang B., Hoober-Burkhardt L., Wang F., Surya Prakash G. K., Narayanan S. R., An inexpensive aqueous flow battery for large-scale electrical energy storage based on water-soluble organic redox couples. J. Electrochem. Soc. 161, A1371–A1380 (2014). [Google Scholar]
- 21.Lin K., et al., A redox-flow battery with an alloxazine-based organic electrolyte. Nat. Energy 1, 16102 (2016). [Google Scholar]
- 22.Kwabi D. G., et al., Alkaline quinone flow battery with long lifetime at PH 12. Joule 2, 1894–1906 (2018). [Google Scholar]
- 23.Darling R. M., Gallagher K. G., Kowalski J. A., Ha S., Brushett F. R., Pathways to low-cost electrochemical energy storage: A comparison of aqueous and nonaqueous flow batteries. Energy Environ. Sci. 7, 3459–3477 (2014). [Google Scholar]
- 24.Brushett F. R., Aziz M. J., Rodby K. E., On lifetime and cost of redox-active organics for aqueous flow batteries. ACS Energy Lett. 5, 879–884 (2020). [Google Scholar]
- 25.Goulet M.-A., Aziz M. J., Flow battery molecular reactant stability determined by symmetric cell cycling methods. J. Electrochem. Soc. 165, A1466–A1477 (2018). [Google Scholar]
- 26.Sankarasubramanian S., Seo J., Mizuno F., Singh N., Prakash J., Elucidating the oxygen reduction reaction kinetics and the origins of the anomalous Tafel behavior at the lithium-oxygen cell cathode. J. Phys. Chem. C 121, 4789–4798 (2017). [Google Scholar]
- 27.Sankarasubramanian S., Ramani V., Dimethyl sulfoxide-based electrolytes for high-current potassium–oxygen batteries. J. Phys. Chem. C 122, 34 (2018). [Google Scholar]
- 28.Li Y., Parrondo J., Sankarasubramanian S., Ramani V., Impact of surface carbonyl- and hydroxyl-group concentrations on electrode kinetics in an all-vanadium redox flow battery. J. Phys. Chem. C 123, 6370–6378 (2019). [Google Scholar]
- 29.Sankarasubramanian S., Kahky J., Ramani V., Tuning anion solvation energetics enhances potassium-oxygen battery performance. Proc. Natl. Acad. Sci. U.S.A. 116, 14899–14904 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Bard A. J., Faulkner L. R., Electrochemical Methods: Fundamentals and Applications (Wiley, ed. 2, 2000). [Google Scholar]
- 31.Bockris J. O., Nagy Z., Symmetry factor and transfer coefficient. A source of confusion in electrode kinetics. J. Chem. Educ. 50, 839 (1973). [Google Scholar]
- 32.Nicholson R. S., Shain I., Theory of stationary electrode polarography: Single scan and cyclic methods applied to reversible, irreversible, and kinetic systems. Anal. Chem. 36, 706–723 (1964). [Google Scholar]
- 33.Nicholson R. S., Theory and application of cyclic voltammetry for measurement of electrode reaction kinetics. Anal. Chem. 37, 1351–1355 (1965). [Google Scholar]
- 34.Marken F., Neudeck A., Bond A. M., Cyclic voltammetry. Electroanal. Methods Guid. to Exp. Appl. 2, 57–106 (2010). [Google Scholar]
- 35.Miller C. C., The Stokes-Einstein law for diffusion in solution. Proc. R. Soc. A Math. Phys. Eng. Sci. 106, 724–749 (1924). [Google Scholar]
- 36.Klingler R. J., Kochi J. K., Electron-transfer kinetics from cyclic voltammetry. Quantitative description of electrochemical reversibility. J. Phys. Chem. 85, 1731–1741 (1981). [Google Scholar]
- 37.Kwabi D. G., Ji Y., Aziz M. J., Electrolyte lifetime in aqueous organic redox flow batteries: A critical review. Chem. Rev. 120, 6467–6489 (2020). [DOI] [PubMed] [Google Scholar]
- 38.Murugavel K., Benzylic viologen dendrimers: A review of their synthesis, properties and applications. Polym. Chem. 5, 5873–5884 (2014). [Google Scholar]
- 39.Bard A. J., Ledwith A., Shine H. J., Formation, properties and reactions of cation radicals in solution. Adv. Phys. Org. Chem. 13, 155–278 (1976). [Google Scholar]
- 40.Zheng D., Li H., Lu B., Xu Z., Chen H., Electrochemical properties of ferrocene adsorbed on multi-walled carbon nanotubes electrode. Thin Solid Films 516, 2151–2157 (2008). [Google Scholar]
- 41.Wu J., Fu Q., Lu B., Li H., Xu Z., Surfactant-associated electrochemical properties of ferrocene adsorbed on a glassy carbon electrode modified with multi-walled carbon nanotubes. Thin Solid Films 518, 3240–3245 (2010). [Google Scholar]
- 42.Lin C., Batchelor-McAuley C., Laborda E., Compton R. G., Tafel-Volmer electrode reactions: The influence of electron-transfer kinetics. J. Phys. Chem. C 119, 22415–22424 (2015). [Google Scholar]
- 43.Henstridge M. C., Laborda E., Rees N. V., Compton R. G., Marcus-Hush-Chidsey theory of electron transfer applied to voltammetry: A review. Electrochim. Acta 84, 12–20 (2012). [Google Scholar]
- 44.Compton R. G., Banks C. E., “Cyclic voltammetry at macroelectrodes” in Understanding Voltammetry (World Scientific Publishing Europe Ltd., 2007), pp. 107–151. [Google Scholar]
- 45.Helfferich F. G., Kinetics of Homogeneous Multistep Reactions (Elsevier, 2001), vol. 38. [Google Scholar]
- 46.Laborda E., Henstridge M. C., Batchelor-McAuley C., Compton R. G., Asymmetric Marcus-Hush theory for voltammetry. Chem. Soc. Rev. 42, 4894–4905 (2013). [DOI] [PubMed] [Google Scholar]
- 47.Savéant J., Costentin C., Elements of Molecular and Biomolecular Electrochemistry (Wiley, 2019). [Google Scholar]
- 48.Bai P., Bazant M. Z., Charge transfer kinetics at the solid-solid interface in porous electrodes. Nat. Commun. 5, 3585 (2014). [DOI] [PubMed] [Google Scholar]
- 49.Atkins P. W., MacDermott A. J., The Born equation and ionic solvation. J. Chem. Educ. 59, 359 (1982). [Google Scholar]
- 50.Fogler H. S., Elements of Chemical Reaction Engineering (Prentice Hall, ed. 4, 2006). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All study data are included in the article and/or SI Appendix.





