ACS Publications. Most Trusted. Most Cited. Most Read
Molecular Level Understanding of the Factors Affecting the Stability of Dimethoxy Benzene Catholyte Candidates from First-Principles Investigations
My Activity

Figure 1Loading Img
  • Open Access
Article

Molecular Level Understanding of the Factors Affecting the Stability of Dimethoxy Benzene Catholyte Candidates from First-Principles Investigations
Click to copy article linkArticle link copied!

View Author Information
† ‡ Joint Center for Energy Storage Research, Materials Science Division, and §Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States
*E-mail: [email protected]. Tel.: 630-252-3536. Fax: 630-252-9555.
Open PDFSupporting Information (1)

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2016, 120, 27, 14531–14538
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.6b04263
Published June 14, 2016

Copyright © 2016 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

First-principles simulations are performed to gain molecular level insights into the factors affecting the stability of seven 1,4-dimethoxybenzene (DMB) derivatives. These molecules are potential catholyte candidates for nonaqueous redox flow battery systems. Computations are performed to predict oxidation potentials in various dielectric mediums, intrinsic-reorganization energies, and structural changes of these representative catholyte molecules during the redox process. In order to understand the stability of the DMB-based radical cations, the thermodynamic feasibility of the following reactions is computed using density functional theory: (a) deprotonation, (b) dimerization, (c) hydrolysis, and (d) demethylation. The computations indicate that radical cations of the 2,3-dimethyl and 2,5-dimethyl derivatives are the most stable among the DMB derivatives considered in this study. In the presence of solvents with high-proton solvating ability (water, DMSO, acetonitrile), degradation of cation radical occurring via deprotonation is the most likely mechanism. In the presence of solvents such as propylene carbonate (PC), demethylation was found to be the most likely reaction that causes degradation of radical cations. From the computed enthalpy of activation (ΔH) for a demethylation reaction in PC, the 2,5-dimethyl DMB cation radical would exhibit better kinetic stability in comparison to the other candidates. This investigation suggests that computational studies of structural properties such as redox potentials, reorganization energies, and the computed reaction energetics (deprotonation and demethylation) of charged species can be used to predict the relative stability of a large set of molecules required for the discovery of novel redox active materials for flow battery applications.

Copyright © 2016 American Chemical Society

1 Introduction

Click to copy section linkSection link copied!

Effective energy storage in organic molecules has a role to play in the area of the grid electricity storage due to the increased production of intermittent renewable electricity. (1-3) Redox couples for such applications need to be chemically and electrochemically (oxidation and reduction) stable to perform thousands of cycles with no significant degradation. In a redox couple, molecules that undergo oxidation and reduction are classified into “catholyte” and “anolyte”, respectively. Prerequisites of ideal candidates for flow battery applications are their adequate electrochemical window, solubility, fast redox kinetics, stability, and cost. Detailed molecular level understandings of all the above properties are crucial to the development of new and improved candidates for grid storage applications.
Recently, the use of redox active organic molecules for energy storage has made exceptional progress in aqueous systems, (4-7) while finding desirable redox flow molecules (catholytes and anolytes) in nonaqueous media was found to be exceptionally challenging. (8-11) Among, many possible catholyte candidates as redox active materials in nonaqueous media, 2,5-di-tert-butyl-1,4-dimethoxybenzene was found to be promising material in the battery operating conditions by Dahn et al. (12, 13) These materials have a reasonable oxidation window (>3.5 V Li/Li+) and stability as radical cations compared to many materials investigated. (14) Similarly, further studies suggest that derivatives of 1,4-dimethoxybenzene (DMB) can be used as stable material as redox shuttle in lithium ion batteries. (14-16) In general, due to the electrochemical window, any DMB derivatives with solubility and stability can be used as catholyte materials in flow batteries. In terms of assessing the feasibility of redox materials, there have been many studies of their redox properties. (17-23) However, the reasons for the differences in stability of organic molecules for flow batteries are poorly understood due to the complexity of the reactivity patterns (24, 25) of the radical cations (RC) in solution. However, information regarding the reactivity of radical cations is critical toward understanding the factors that control stability. Therefore, in this investigation we utilize the predictive ability of density functional theory (DFT) to gain insights into factors controlling the stability of seven catholyte candidates based on DMB derivatives. The structures of the seven molecules are shown in Figure 1.

Figure 1

Figure 1. Schematic of dimethoxybenzene (DMB) derivatives considered in this study.

A critical bottleneck for understanding the stability is the complexity of the radical cations due to their many intrinsic and extrinsic properties. Intrinsic properties of radical cations include the following: (i) delocalization of spin and charge, (ii) release of strain, (iii) stereo electronic factors, and (iv) steric hindrance. Extrinsic properties include the following: (v) solvent medium, (vi) anode materials, and (vii) electrolytes. A detailed understanding of these intrinsic and extrinsic effects requires comprehensive experimental and theoretical studies. In this paper we present a theoretical investigation focused on understanding key features that affect the stability of radical cations (RC) formed during the oxidation of the catholytes. These features and associated quantum chemical descriptors that can be computed are schematically shown in Figure 2. We present results for the electrochemical windows, structural distortion, and reactivity of radical cations of the seven DMB derivatives. The oxidation potential and reorganization energy are hypothesized as the descriptors for electrochemical window and structural distortion, respectively. Additionally, computed free energies of reactions such as deprotonation, dimerization, hydrolysis, and demethylation of cation radicals are hypothesized as the descriptors for their reactivity.

Figure 2

Figure 2. Key properties controlling the stability of redox active species and the quantum chemical descriptors that can assess the properties are schematically shown. Note: G, E, and λ represent Gibbs free energy, redox potential, and reorganization energy, respectively.

Details of the computational approach employed in this investigation are presented in the next section. In Results and Discussion, computational studies are presented that rank the stability of selected catholyte molecules using structural and energetic parameters.

2 Computational Details

Click to copy section linkSection link copied!

All computations were performed using the Gaussian 09 software. (26) The B3LYP/6-31G(2df,p) level of theory was used to compute the structure, electronic energy, vibrational frequencies, and free energy corrections of all species. Solvation free energies were computed with the SMD solvation model by performing single point energy evaluations with a specific dielectric medium: diethyl ether (ε = 4.24), acetone (ε = 20.49), dimethyl sulfoxide (ε = 46.83), or water (ε = 78.36). Reorganization energies were computed (section 3.1.2) from the structures of the molecules optimized in the diethyl ether solvent medium. We note that explicit interactions of solvent molecules may be essential to represent charged states of the molecules/ions are not included in this investigation due to the computational complexity, which is beyond the scope of this manuscript.
The oxidation potentials (Eox w.r.t. Li/Li+) were computed from the computation of Gibbs free energy change (ΔGox, eV) at 298 K in the solution (dielectric) for the removal of an electron from the species of interest, using the following equation:where F is the Faraday constant (in eV) and n is the number of electrons involved in the oxidation process. The addition of the constant “–1.24 V” is required to convert the free energy changes to oxidation potential (Li/Li+ reference electrode), a commonly used convention to compute the redox potentials in solution. (27, 28) The change in energy of electrons when going from vacuum to nonaqueous solution is treated as zero, similar to what has been used by others. (29) Further details regarding the computation of redox potential can be found elsewhere. (29-34)
Solvation energy contributions of a species are added to the gas phase enthalpies and free energies to approximate enthalpies (Hsoln = Hgas phase + ΔGsolv) and free energies (Gsoln = Ggas phase + ΔGsolv) in solution. The demethylation (C–O bond breaking) reaction barriers presented in section 3.2.2 are apparent enthalpy barriers (see SI for activation free energy barriers), computed as the difference between the enthalpy of the transition state structure (H) and the sum of enthalpies of reactants in solution at 298 K. Here we used the solution phase enthalpy approximation because the main free energy contributions (GCDS: cavitation, dispersion, and solvent structure terms) cancel out (or are negligible) when computing the apparent barriers (H(TS)–H(reactants)). Therefore, the dominant solvation contributions are electronic, nuclear, and polarization terms (GENP), which are included in the computation of solution enthalpies. A similar approach was successfully employed elsewhere. (35) The B3LYP/6-31G(2df,p) level of theory is used to compute the transition state structures in the gas phase and, subsequently, a single point solvation energy calculation is performed using the SMD solvation model (diethyl ether and water) at the same level of theory. Thus, the enthalpy barriers (ΔH) presented in section 3.2.2 are the sum of the gas phase enthalpy barrier and the solvation energy (ESMD):

3 Results and Discussion

Click to copy section linkSection link copied!

In this section, computation of various structural and reactivity related properties (Figure 2) of the seven DMB-based molecules are described. The structure related properties, electrochemical windows and structural deviation upon oxidation, are discussed in section 3.1. The reactivity related properties such as thermochemical data for deprotonation, dimerization, hydrolysis, and demethylation reactions of radical cations and their relevance are discussed in section 3.2.

3.1 Structure

Three properties related to structural stability of the dimethoxybenzene (DMB) derivatives were computed at the B3LYP/6-31G(2df,p) level of theory. These properties are (i) oxidation potential, (ii) reorganization energy (λ), and (iii) the Kabsch root-mean-square deviation (RMSD) between the optimized coordinates of the oxidized and neutral state of the molecules.

3.1.1 Simulated Oxidation Potentials (Eox)

The electrochemical windows of the seven DMB derivatives in various dielectric media including diethyl ether (DEE; ε = 4.24), acetone (ε = 20.49), dimethyl sulfoxide (DMSO; ε = 46.83), and water (ε = 78.36) were computed and are presented in Table 1. In a low dielectric medium such as in diethyl ether, the oxidation potentials are in the range of 3.9–4.5 V for the DMB candidate molecules. Among the molecules in diethyl ether medium, the 25-DMB and 23-DMB exhibit lower oxidation potentials (3.9 V each), while 235-DMB exhibits a high oxidation potential (4.5 V). At high dielectrics such as DMSO (ε = 46.83) and water (ε = 78.36), the computed oxidation potentials are on the order of 3.5–4.1 V. This slight reduction in oxidation potential in a high dielectric medium compared to low dielectric conditions is expected due to the increased stabilization of the charged state by the high dielectric medium compared to the neutral state.
Table 1. Computed Oxidation Potentials (V), Reorganization Energies (λtotal, λox), and Root Mean Squared Deviation (RMSD) between the Optimized Cartesian Coordinates of Neutral and Oxidized Species of DMB Derivatives (see Figure 1)a
 computed oxidation potential (V, Li/Li+)reorganization energy (kcal/mol) 
speciesEDEEEacetoneEDMSOEH2OλoxλtotalKabsch RMSD
DMBa4.23.83.73.85.611.30.04
2-DMB4.13.73.73.75.511.20.04
23-DMB3.93.63.53.615.619.70.63
25-DMB3.93.63.53.65.411.10.04
26-DMB4.13.83.73.710.921.30.59
235-DMB4.54.14.14.19.120.40.44
2356-DMBb4.33.93.93.914.631.90.52
a

Note: The total reorganization energy (λtotal) is defined as the sum of the reorganization energy during the oxidation (λox) and reduction (λred). Abbreviations: DEE, diethyl ether; DMSO, dimethyl sulfoxide; H2O, water. Notes: (a) experimentally (12) measured oxidation potentials in propylene carbonate are 3.9 and 4.1 V for DMB (a) and 2356-DMB (b) molecules, respectively.

3.1.2 Reorganization Energy (λ)

The reorganization energy (λ) of a molecule during a redox process is computed using the energies and geometries of the neutral and oxidized states. The connection between the nuclear configuration and ground state energies of a given molecule and its cation radical is schematically shown in Figure 3. The total reorganization energy (λtotal) is defined as the sum of the reorganization energy during the oxidation (λox) and reduction (λred). We note that zero-point energy correction is not considered in the reorganization energy computations. The λox is the energy difference between the vertical detachment energy (VDE) and adiabatic detachment energy (ADE). The (λred) is the energy difference between ADE and vertical electron affinity (VEA). The absolute values of reorganization energy are used here and are tabulated in Table 1. Lower reorganization energy suggests that the molecular system achieves energy closer to the minimum energy upon electron addition or removal. For oxidation processes, based on the computed λox presented in Table 1, the 25-DMB has the lowest and the 23-DMB has the highest requirement of reorganization energy. We note that, the 2356-DMB molecule requires highest total reorganization energy as expected due to six substituent groups in the benzene ring. In general, from Table 1, the unsubstituted DMB, 2-DMB, and the 25-DMB have lower values for λox and λtotal.

Figure 3

Figure 3. Schematic of the computation of reorganization energy during oxidation (λox) and reduction (λred) process from the computed energies of a molecule in its neutral and singly positive charge state. Total reorganization energy ((λtotal) is the sum of reorganization energy required during the oxidation (λox) and reduction (λred) process.

3.1.3 Root Mean Squared Deviation (RMSD) between the Geometries

Another measure of the structural deviation during the redox processes is the RMSD between the optimized Cartesian coordinates of neutral and oxidized species of the DMB derivatives. From the neutral and unipositive DFT optimized (xyz) structures of molecules (all optimized structures are given in the Supporting Information, Figure S1 and Table S1); the root-mean-square deviation (RMSD) was computed using the Kabsch algorthm. (36) The computed RMSD is also shown in Table 1. Based on the computations, the 25-DMB exhibits a lowest RMSD for geometry change (0.04) and also the lowest reorganization energy, suggesting stabilization over the rest of the candidate molecules. Two other derivatives (DMB and 2-DMB) have similarly low RMSDs during oxidation consistent with the prediction of reorganization energy.
To compare the oxidation potentials (Eox), reorganization energies (λox), and RMSD during oxidation of the candidate molecules, these three properties are plotted in Figure 4. From the figure it is evident that the 25-DMB exhibits low values for all three properties. Having lower oxidation potentials (Eox) in a given series of molecules is an indicator of relative stability of cation radicals. This is based on the assumption that these cation radicals are less reactive toward typical deprotonation reactions. Methylated DMB derivatives have similar C–H bond dissociation energies; therefore, the acidity of the corresponding cation radicals increase with the oxidation potential of neutral species. (37) In this aspect, the 25-DMB and 23-DMB are less acidic (lower computed oxidation potential, see Table 1) compared to the rest of the candidates. In addition by incorporating other structural features such as reorganization energy (λ) or RMSD during the oxidation, it is evident that 25-DMB can be considered to be the most stable catholyte material from the seven DMB derivatives considered. In general, plots similar to that of Figure 4 can be obtained for a series of molecules to assess the relative stability potential redox active materials. Further understanding of the fate of the radical cation is also essential to understand the relative stability of radical cations and the overall stability of materials. We consider this in section 3.2).

Figure 4

Figure 4. Comparison of computed oxidation potentials (Eoxn V unit w.r.t. Li/Li+ ref electrode), reorganization energy during oxidation (λox), and root-mean-square deviation (RMSD) between the optimized xyz coordinates of neutral and oxidized states of various DMB derivatives (x-axis). Note: The data associated with the plot is given in Table 1.

3.2 Reactivity

To assess the reactivity of cation radicals, it is essential to understand the thermodynamic driving forces and kinetic feasibilities of likely reactions. In subsection 3.2.1, the computed thermochemistry of selected probe reactions of radical cations (deprotonation, dimerization, hydrolysis, and demethylation) is discussed. In subsection 3.2.2, the kinetic barriers of demethylation reactions are evaluated and discussed.

3.2.1 Thermodynamic Feasibility of Chemical Reactions

The thermodynamic feasibility (ΔG) of deprotonation (rxn1), demethylation (rxn2), hydrolysis (rxn3), and dimerization (rxn4) reactions were computed for the radical cations of (RCs) the candidates. These four reactions are schematically shown for DMB in Figure 5. The deprotonation (rxn 1) of the cation radical in water is chosen as the probe reaction, where deprotonation results into the formation of neutral radical and a solvated proton. This reaction is considered due to the available experimental data of free energy of solvation of the proton in aqueous medium ΔGproton solvation = −265.9 kcal/mol, (38) which is needed to understand the thermochemistry of protonation and deprotonation. (39) Deprotonation of DMB derivatives could occur from the methoxy (−OCH3) or from the methyl (−CH3) group. Note that the deprotonation from the benzene ring is unlikely due to the largely endothermic nature (∼+40 kcal/mol) of the reaction. Similar to deprotonation, demethylation (rxn 2) is another way of eliminating the positive charge of the radical cations. To address rxn 2, we have computed the thermochemistry for methyl transfer between the radical cation (RC) and propylene carbonate (PC), a probe molecule, as shown in Figure 5. In addition to the deprotonation and demethylation reactions, a hydrolysis reaction (rxn 3), a probe reaction for solvolysis, is considered. In this reaction the thermochemistry of etheric cleavage by a water molecule to form CH3OH and hydroxyl radical cations (see Figure 5, rxn3) was studied. Finally, dimerization (rxn4) of the cation radical was also considered as a probe reaction. The computed Gibbs free energies of rxns1–4 are presented in Table 2. Using the Gibbs free energy of deprotonation (ΔG(H+)aq) and Gibbs free energy of hydrolysis (ΔGHOH), the pKa and KHOH equilibrium constants, respectively, were also computed and presented in Table 2.

Figure 5

Figure 5. Schematic of selected reactions (rxn1–4) of 1,4-dimethoxybenzene (DMB) radical cation. Abbreviation: PC, propylene carbonate molecule. All reactions are modeled in aqueous dielectric medium. In rxn1, the cation radical reacts with water to form solvated proton and neutral radical. In rxn4, two cation radicals interact with water to form a dimer and two solvated protons.

Table 2. Computed Thermochemical Data (in kcal/mol) of DMB Cation Radicals at the B3LYP/6-31G(2df,p) Level of Theorya
 deprotonation from “–OCH3” groupdeprotonation from “–CH3” group 
radical cation (RC)ΔG(H+)aqbpKacΔGdimeraqdΔG(H+)aqepKacΔGdimeraqdgΔG(CH3+)PCΔGHOHhKHOHi
DMB28.37.3–12.0NANANA14.00.61.0 × 10–01
2-DMB29.97.7–8.321.15.4–1.113.70.16.5 × 10–01
23-DMB31.38.1–5.022.75.84.214.9–0.22.3 × 1000
25-DMB32.48.3–3.323.46.03.615.40.25.0 × 10–01
26-DMB27.27.0–9.320.85.4–3.610.0–4.53.6 × 1007
235-DMB19.35.0–27.0f10.82.8–19.99.4–6.94.2 × 1011
2356-DMB22.75.9–20.414.93.8–10.75.9–7.18.6 × 1011
a

Computed data includes deprotonation free energies (ΔG(H+)aq), dimerization energies (ΔGdimeraq), and demethylation energies (ΔG(CH3+)) of DMB cation radicals.

b

Free energy change during the deprotonation reaction (rxn1) from the methoxy group of the DMB derivative using the experimental Gibbs free energy of proton in aqueous solution (−265.9 kcal/mol).

c

pKa = ΔG/2.303RT, where R is the gas constant and T is the absolute temperature.

d

ΔGdimer is the free energy change during the dimerization (rxn4) of two cation radicals to form neutral dimer, where two protons are solvated by an aqueous medium.

e

The computed free energy change for deprotonation (rxn1) from the methyl group to the aqueous medium using the Gibbs free energy of a proton in aqueous solution: −265.9 kcal/mol. (38)

f

For 235-DMB, the deprotonation from the 3-methyl position (ΔG(H+)aq = 10.8 kcal/mol) is thermodynamically more preferred than from the 2-methyl (ΔG(H+)aq = 12.3 kcal/mol) and from the 5-methyl (ΔG(H+)aq = 11.7 kcal/mol) position.

g

ΔG = G(RC) + G(PC) → G(PC–CH3+) + G(RC-demethylated), see rxn2 in Figure 5.

h

ΔG = G(RC) + G(H2O) → G(CH3OH) + G(RC-H) radical, barrier is not computed.

i

The equilibrium constant for hydrolysis, KHOH = exp(−ΔGHOH/2.303RT).

Based on the computed data presented in Table 2, the following important points can be drawn:
1.

Deprotonation of DMB radical cations are endothermic processes in aqueous solution (10–32 kcal/mol). Deprotonation from a methyl group (10–23 kcal/mol) is thermodynamically more likely than from the methoxy groups (19–32 kcal/mol) for all DMB derivatives.

2.

Dimerization of cation radicals are generally thermodynamically downhill, and it is likely that the rate-determining deprotonation is the first step. Dimerization of neutral radical formed at the methoxy group (−OCH2−) is thermodynamically more favorable than that formed at the methyl (−CH2– group). See Figure S3 of the Supporting Information for details.

3.

Demethylation reactions of RCs by a PC molecule in aqueous medium are endoergic reactions (5–16 kcal/mol). This is the most likely reaction based on the free energy of the reaction.

4.

Similar to demethylation, hydrolysis can also cleave the etheric C–O bond. Based on the computations in Table 2, hydrolysis of DMB analogues with steric constraints (due to methyl groups) are thermodynamically favorable.

Additionally, deprotonation energetics are likely to be more thermodynamically uphill in solvents such as methanol (ΔGproton solvation = −263.5 kcal/mol (38)) or acetonitrile (ΔGproton solvation = −260.2 kcal/mol (38)) than in water medium, due to the reduced proton solvation energy of methanol and acetonitrile compared to the aqueous medium (ΔGproton solvation = −265.9 kcal/mol). In terms of the reactivity of cation radicals, deprotonation and demethylation are the most likely reactions since both of them are ideal for the radical cations to lose its charge to the solvent medium. The dimerization reaction is a consequence of the deprotonation and hydrolysis reaction (or solvolysis) is one of the routes for the demethylation reaction. In general, based on the computed reaction free energies of demethylation reaction, most likely based on the free energies, the 25-DMB radical cations would exhibit the least reactivity in the solution, indicating better relative stability as catholyte materials.

3.2.2 Demethylation: Computed Activation Barriers

Table 3. Computed Enthalpy of Activation (kcal/mol) for Demethylation of DMB Radical Cations by a Propylene Carbonate Molecule at the B3LYP/6-31G(2df,p) Level of Theory in Water and Diethyl Ether Solvent Dielectrica
 water dielectricdiethyl ether dielectric
radical cation (RC)ΔH(CH3+)PCKb (s–1)T1/2c (h)ΔH(CH3+)PCbKc (s–1)T1/2d (h)
DMB26.71.6 × 10–71.2 × 10323.63.0 × 10–56.4
2-DMB26.52.5 × 10–77.9 × 10223.82.6 × 10–58.5
23-DMB27.26.8 × 10–82.8 × 10324.74.9 × 10–63.9 × 101
25-DMB27.92.1 × 10–89.1 × 10325.69.9 × 10–71.9 × 102
26-DMB23.82.3 × 10–58.521.02.5 × 10–37.7 × 10–2
256-DMB16.54.84.0 × 10–513.94.4 × 1024.4 × 10–7
2356-DMB23.72.0 × 10–59.621.51.0 × 10–31.9 × 10–1
a

Using the enthalpy of activation in solution, the rate constant (K) and half-life assuming first order kinetics are shown.

b

ΔG(CH3+)PC is gven in the Table S2 of the Supporting Information;

c

K = (KbT/hc)exp(−ΔG/RT).

d

T1/2 = ln(2)/K (h–1).

In addition to the relative stability predictions based on the computed free energy changes associated with the deprotonation/demethylation reactions, the activation barriers for demethylation of DMB analogues were also computed. In general, calculation of activation barriers using the demethylation model can provide qualitative trends for describing the initial decomposition reactions of cation radicals. Demethylation of DMB cation analogues by a propylene carbonate molecule in water dielectric medium is used as a model to compute the activation barriers. In Figure 6, a computed enthalpy profile for the demethylation reaction of 25-DMB is shown, where the computed apparent activation enthalpy (ΔH) is ∼28 kcal/mol. In the energy profile (Figure 6), first a weakly bound complex between 25-DMB cation radical and a PC molecule is formed. Subsequently, the complex undergoes C–Omethyl bond cleavage via the transition state shown in the Figure 6. Complete transfer of methyl cation to the PC results in the formation of a 25-DMB demethylated radical. Note that the demethylation reactions are endothermic in nature, as explained in the previous section (see Table 2). In Table 3, the computed apparent activation barriers (ΔH) for all radical cations in aqueous and diethyl ether solvent model are presented. A comparison of computed activation enthalpies and the free energy changes associated with the demethylation of DMB cation radicals by propylene carbonate is schematically shown in Figure 7. Using this activation enthalpy in solution, we have approximated the rate constant (K, s–1) using the Eyring equation (see Table 3). Additionally, using first order kinetics approximation, the half-life of the reactions (T1/2, h) is also computed, also shown in Table 3. Based on the computations, the activation barriers for the demethylation are smaller in diethyl ether compared to water. Most importantly, the computed activation barriers are in the order: 25-DMB > 23-DMB > 2-DMB > DMB > 2356-DMB > 26-DMB > 256-DMB in ether dielectric medium. The computed half-life of the demethylation reaction suggest that both 25-DMB and 23-DMB would require approximately hundreds of hours to decompose via demethylation reaction, while the rest of the cation radicals require tens of hours or less. This is consistent with the conclusion above based on the computed thermochemistry that these two would be the most stable.

Figure 6

Figure 6. Computed solution phase enthalpy profile (ΔHsoln) of demethylation of the 25-DMB cation radical by a propylene carbonate molecule. Computed apparent enthalpy of activation (ΔH) of all DMB candidates are given in Table 3.

Figure 7

Figure 7. Comparison of the computed free energies changes (ΔG(CH3+)PC) and activation enthalpies (ΔH(CH3+)PC) required for the demethylation of DMB cation radicals by propylene carbonate. The data associated with (ΔG(CH3+)PC) and (ΔH(CH3+)PC) are given in Tables 2 and 3, respectively.

4 Summary

Click to copy section linkSection link copied!

Stable, electrochemically active and soluble organic materials are required for redox flow energy storage applications. Understanding the stability of organic materials in the battery operating conditions is a complex problem, but essential to provide guidelines for materials discovery. In this study, first-principles simulations are performed to gain molecular level insights into the factors affecting the stability of seven 1,4-dimethoxybenzene (DMB) derivatives. These molecules are potential catholyte candidates for nonaqueous redox flow battery systems. Computations are performed to predict oxidation potentials in various dielectric mediums, intrinsic-reorganization energies, and structural changes of these representative catholytes during the redox process. In order to understand the stability of the DMB-based radical cations, the thermodynamic feasibility of the following reactions is computed using density functional theory: (a) deprotonation, (b) dimerization, (c) hydrolysis, and (d) demethylation.
The computations indicate that radical cations of the 2,3-dimethyl and 2,5-dimethyl derivatives are the most stable among the DMB derivatives considered in this study. In the presence of solvents with high-proton solvating ability (water, DMSO, acetonitrile), degradation of the cation radical likely occurs via deprotonation. In the presence of solvents such as propylene carbonate (PC), demethylation was found to be the most likely reaction that causes degradation of radical cations. From the computed enthalpy of activation (ΔH) for a demethylation reaction in PC, the 2,5-dimethyl DMB cation radical would exhibit better kinetic stability in comparison with the other candidates.
This investigation suggests that computational studies of structural properties such as redox potentials, reorganization energies, and the computed reaction energetics (deprotonation and demethylation) of charged species could be used to predict the relative stabilities of large sets of molecules required for the discovery of novel redox active materials for flow battery applications. First, computationally less demanding, reorganization energies or RMSD between structures can be used as a descriptor for relative stability. Second, a reactivity indicator from computed free energy change of the most likely reaction can be used to screen a large set of molecules to rank the stability trends. Finally, as part of the deep dive study, for a small subset of radical cations kinetic feasibility of the most likely reaction can be computed to assess the reactivity and lifetime. We also anticipate that these first-principles-based computations can be used in the future to derive quantitative relationships to predict the stability of materials for redox flow applications.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.6b04263.

  • Optimized structures of DMB derivatives are shown in Figure S1. Figure 2 suggests most important structural parameters and these parameters are tabulated in Table S1. Figure S3 presents the computed deprotonation energies (rxn 1 of Figure 5) and dimerization energies (rxn 4 of the Figure 5) of radical cations at the B3LYP/6-31G(2df,p) level of theory, and Table S2 shows computed enthalpies and free energies of activation required for the demethylation reactions in water and diethyl ether dielectric media (PDF).

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
    • Rajeev S. Assary - †Joint Center for Energy Storage Research, ‡Materials Science Division, and §Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States Email: [email protected]
  • Authors
    • Lu Zhang - †Joint Center for Energy Storage Research, ‡Materials Science Division, and §Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States
    • Jinhua Huang - †Joint Center for Energy Storage Research, ‡Materials Science Division, and §Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States
    • Larry A. Curtiss - †Joint Center for Energy Storage Research, ‡Materials Science Division, and §Chemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States
  • Notes
    The authors declare no competing financial interest.

Acknowledgment

Click to copy section linkSection link copied!

This work was supported as part of the Joint Center for Energy Storage Research, an Energy Innovation Hub funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences. We gratefully acknowledge the computing resources provided on “Blues”, a 320-node computing cluster operated by the Laboratory Computing Resource Center at Argonne National Laboratory.

References

Click to copy section linkSection link copied!

This article references 39 other publications.

  1. 1
    Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical Energy Storage for the Grid: A Battery of Choices Science 2011, 334, 928 935 DOI: 10.1126/science.1212741
  2. 2
    Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi, D.; Lemmon, J. P.; Liu, J. Electrochemical Energy Storage for Green Grid Chem. Rev. 2011, 111, 3577 3613 DOI: 10.1021/cr100290v
  3. 3
    Larcher, D.; Tarascon, J. M. Towards Greener and more Sustainable Batteries for Electrical Energy storage Nat. Chem. 2014, 7, 19 29 DOI: 10.1038/nchem.2085
  4. 4
    Huskinson, B.; Marshak, M. P.; Suh, C.; Er, S.; Gerhardt, M. R.; Galvin, C. J.; Chen, X.; Aspuru-Guzik, A.; Gordon, R. G.; Aziz, M. J. A Metal-free Organic-Inorganic Aqueous Flow Battery Nature 2014, 505, 195 198 DOI: 10.1038/nature12909
  5. 5
    Lin, K.; Chen, Q.; Gerhardt, M. R.; Tong, L.; Kim, S. B.; Eisenach, L.; Valle, A. W.; Hardee, D.; Gordon, R. G.; Aziz, M. J.; Marshak, M. P. Alkaline Quinone Flow Battery Science 2015, 349, 1529 1532 DOI: 10.1126/science.aab3033
  6. 6
    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. 2014, 161, A1371 A1380 DOI: 10.1149/2.1001409jes
  7. 7
    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. 2016, 6, 1501449 1501456 DOI: 10.1002/aenm.201501449
  8. 8
    Wei, X.; Cosimbescu, L.; Xu, W.; Hu, J. Z.; Vijayakumar, M.; Feng, J.; Hu, M. Y.; Deng, X.; Xiao, J.; Liu, J.; Sprenkle, V.; Wang, W. Towards High-Performance Nonaqueous Redox Flow Electrolyte Via Ionic Modification of Active Species Adv. Energy Mater. 2015, 5, 1400678 1400685 DOI: 10.1002/aenm.201400678
  9. 9
    Brushett, F. R.; Vaughey, J. T.; Jansen, A. N. An All-Organic Non-aqueous Lithium-Ion Redox Flow Battery Adv. Energy Mater. 2012, 2, 1390 1396 DOI: 10.1002/aenm.201200322
  10. 10
    Gong, K.; Fang, Q.; Gu, S.; Li, S. F. Y.; Yan, Y. Nonaqueous Redox-flow Batteries: Organic Solvents, Supporting electrolytes, and Redox pairs Energy Environ. Sci. 2015, 8, 3515 3530 DOI: 10.1039/C5EE02341F
  11. 11
    Sevov, C. S.; Brooner, R. E. M.; Chénard, E.; Assary, R. S.; Moore, J. S.; Rodríguez-López, J.; Sanford, M. S. Evolutionary Design of Low Molecular Weight Organic Anolyte Materials for Applications in Nonaqueous Redox Flow Batteries J. Am. Chem. Soc. 2015, 137, 14465 14472 DOI: 10.1021/jacs.5b09572
  12. 12
    Buhrmester, C.; Chen, J.; Moshurchak, L.; Jiang, J.; Wang, R. L.; Dahn, J. R. Studies of Aromatic Redox Shuttle Additives for LiFePO4-Based Li-Ion Cells J. Electrochem. Soc. 2005, 152, A2390 A2399 DOI: 10.1149/1.2098265
  13. 13
    Chen, J.; Buhrmester, C.; Dahn, J. R. Chemical Overcharge and Overdischarge Protection for Lithium-Ion Batteries Electrochem. Solid-State Lett. 2005, 8, A59 A62 DOI: 10.1149/1.1836119
  14. 14
    Zhang, L.; Zhang, Z.; Redfern, P. C.; Curtiss, L. A.; Amine, K. Molecular Engineering towards Safer Lithium-ion Batteries: A highly Stable and Compatible Redox shuttle for Overcharge protection Energy Environ. Sci. 2012, 5, 8204 8207 DOI: 10.1039/c2ee21977h
  15. 15
    Zhang, L.; Zhang, Z.; Wu, H.; Amine, K. Novel redox shuttle Additive for High-Voltage Cathode Materials Energy Environ. Sci. 2011, 4, 2858 2862 DOI: 10.1039/c0ee00733a
  16. 16
    Zhang, Z.; Zhang, L.; Schlueter, J. A.; Redfern, P. C.; Curtiss, L.; Amine, K. Understanding the Redox shuttle stability of 3,5-di-tert-butyl-1,2-dimethoxybenzene for Overcharge Protection of Lithium-ion Batteries J. Power Sources 2010, 195, 4957 4962 DOI: 10.1016/j.jpowsour.2010.02.075
  17. 17
    Speelman, A. L.; Gillmore, J. G. Efficient Computational Methods for Accurately Predicting Reduction Potentials of Organic Molecules J. Phys. Chem. A 2008, 112, 5684 5690 DOI: 10.1021/jp800782e
  18. 18
    Er, S.; Suh, C.; Marshak, M. P.; Aspuru-Guzik, A. Computational Design of Molecules for an all-Quinone Redox Flow Battery Chem. Sci. 2015, 6, 885 893 DOI: 10.1039/C4SC03030C
  19. 19
    Assary, R. S.; Brushett, F. R.; Curtiss, L. A. Reduction Potential Predictions of some Aromatic Nitrogen-Containing Molecules RSC Adv. 2014, 4, 57442 57451 DOI: 10.1039/C4RA08563A
  20. 20
    Bachman, J. E.; Curtiss, L. A.; Assary, R. S. Investigation of the Redox Chemistry of Anthraquinone Derivatives Using Density Functional Theory J. Phys. Chem. A 2014, 118, 8852 8860 DOI: 10.1021/jp5060777
  21. 21
    Hernández-Burgos, K.; Burkhardt, S. E.; Rodríguez-Calero, G. G.; Hennig, R. G.; Abruña, H. D. Theoretical Studies of Carbonyl-Based Organic Molecules for Energy Storage Applications: The Heteroatom and Substituent Effect J. Phys. Chem. C 2014, 118, 6046 6051 DOI: 10.1021/jp4117613
  22. 22
    Pineda Flores, S. D.; Martin-Noble, G. C.; Phillips, R. L.; Schrier, J. Bio-Inspired Electroactive Organic Molecules for Aqueous Redox Flow Batteries. 1. Thiophenoquinones J. Phys. Chem. C 2015, 119, 21800 21809 DOI: 10.1021/acs.jpcc.5b05346
  23. 23
    Karlsson, C.; Jämstorp, E.; Strømme, M.; Sjödin, M. Computational electrochemistry study of 16 isoindole-4, 7-diones as candidates for organic cathode materials J. Phys. Chem. C 2012, 116, 3793 3801 DOI: 10.1021/jp211851f
  24. 24
    Coote, M. L.; Lin, C. Y.; Beckwith, A. L. J.; Zavitsas, A. A. A comparison of methods for measuring relative radical stabilities of carbon-centred radicals Phys. Chem. Chem. Phys. 2010, 12, 9597 9610 DOI: 10.1039/c003880f
  25. 25
    Cheng, L.; Assary, R. S.; Qu, X.; Jain, A.; Ong, S. P.; Rajput, N. N.; Persson, K.; Curtiss, L. A. Accelerating Electrolyte Discovery for Energy Storage with High-Throughput Screening J. Phys. Chem. Lett. 2015, 6, 283 291 DOI: 10.1021/jz502319n
  26. 26
    Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Gaussian 09; Gaussian, Inc.: Wallingford, CT, U.S.A., 2009.
  27. 27
    Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Aqueous Solvation Free Energies of Ions and Ion-Water Clusters Based on an Accurate Value for the Absolute Aqueous Solvation Free Energy of the Proton J. Phys. Chem. B 2006, 110, 16066 16081 DOI: 10.1021/jp063552y
  28. 28
    Bhattacharyya, S.; Stankovich, M. T.; Truhlar, D. G.; Gao, J. Combined Quantum Mechanical and Molecular Mechanical Simulations of One- and Two-Electron Reduction Potentials of Flavin Cofactor in Water, Medium-Chain Acyl-CoA Dehydrogenase, and Cholesterol Oxidase J. Phys. Chem. A 2007, 111, 5729 5742 DOI: 10.1021/jp071526+
  29. 29
    Guerard, J. J.; Arey, J. S. Critical Evaluation of Implicit Solvent Models for Predicting Aqueous Oxidation Potentials of Neutral Organic Compounds J. Chem. Theory Comput. 2013, 9, 5046 5058 DOI: 10.1021/ct4004433
  30. 30
    Moens, J.; Geerlings, P.; Roos, G. A Conceptual DFT Approach for the Evaluation and Interpretation of Redox Potentials Chem. - Eur. J. 2007, 13, 8174 8184 DOI: 10.1002/chem.200601896
  31. 31
    Borodin, O.; Behl, W.; Jow, T. R. Oxidative Stability and Initial Decomposition Reactions of Carbonate, Sulfone, and Alkyl Phosphate-Based Electrolytes J. Phys. Chem. C 2013, 117, 8661 8682 DOI: 10.1021/jp400527c
  32. 32
    Vollmer, J. M.; Curtiss, L. A.; Vissers, D. R.; Amine, K. Reduction Mechanisms of Ethylene, Propylene, and Vinylethylene Carbonates: A Quantum Chemical Study J. Electrochem. Soc. 2004, 151, A178 A183 DOI: 10.1149/1.1633765
  33. 33
    Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Single-Ion Solvation Free Energies and the Normal Hydrogen Electrode Potential in Methanol, Acetonitrile, and Dimethyl Sulfoxide J. Phys. Chem. B 2007, 111, 408 422 DOI: 10.1021/jp065403l
  34. 34
    Ho, J.; Coote, M. L. First-principles prediction of acidities in the gas and solution phase Wiley Interdisc. Rev.: Comput. Mol. Sci. 2011, 1, 649 660 DOI: 10.1002/wcms.43
  35. 35
    Assary, R. S.; Curtiss, L. A.; Moore, J. S. Toward a Molecular Understanding of Energetics in Li–S Batteries Using Nonaqueous Electrolytes: A High-Level Quantum Chemical Study J. Phys. Chem. C 2014, 118, 11545 11558 DOI: 10.1021/jp5015466
  36. 36
    Kabsch, W. A solution for the Best Rotation to Relate Two Sets of Vectors Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 922 923 DOI: 10.1107/S0567739476001873
  37. 37
    Schmittel, M.; Burghart, A. Understanding Reactivity Patterns of Radical Cations Angew. Chem., Int. Ed. Engl. 1997, 36, 2550 2589 DOI: 10.1002/anie.199725501
  38. 38
    Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Single-Ion Solvation Free Energies and the Normal Hydrogen Electrode Potential in Methanol, Acetonitrile, and Dimethyl Sulfoxide J. Phys. Chem. B 2007, 111, 408 422 DOI: 10.1021/jp065403l
  39. 39
    Assary, R. S.; Kim, T.; Low, J. J.; Greeley, J.; Curtiss, L. A. Glucose and Fructose to Platform Chemicals: Understanding the Thermodynamic Landscapes of Acid-Catalysed Reactions Using High-Level Ab Initio Methods Phys. Chem. Chem. Phys. 2012, 14, 16603 16611 DOI: 10.1039/c2cp41842h

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 36 publications.

  1. Jaehyun Park, Farshud Sorourifar, Madhav R. Muthyala, Abigail M. Houser, Madison Tuttle, Joel A. Paulson, Shiyu Zhang. Zero-Shot Discovery of High-Performance, Low-Cost Organic Battery Materials Using Machine Learning. Journal of the American Chemical Society 2024, 146 (45) , 31230-31239. https://doi.org/10.1021/jacs.4c11663
  2. Akash Jain, Ilya A. Shkrob, Hieu A. Doan, Keir Adams, Jeffrey S. Moore, Rajeev S. Assary. Active Learning Guided Computational Discovery of Plant-Based Redoxmers for Organic Nonaqueous Redox Flow Batteries. ACS Applied Materials & Interfaces 2023, 15 (50) , 58309-58319. https://doi.org/10.1021/acsami.3c11741
  3. Kuluni Perera, Wenting Wu, Kaelon A. Jenkins, Michael F. Espenship, Matthias Zeller, Liyan You, Mustafa Ahmed, Kai Lang, Guangchao Liu, Jagrity Chaudhary, Ashkan Abtahi, Dylan Forbes, Julia Laskin, Brett M. Savoie, Jianguo Mei. Degradation Pathways of Conjugated Radical Cations. Chemistry of Materials 2023, 35 (21) , 9135-9149. https://doi.org/10.1021/acs.chemmater.3c01854
  4. Madison R. Tuttle, Emma M. Brackman, Farshud Sorourifar, Joel Paulson, Shiyu Zhang. Predicting the Solubility of Organic Energy Storage Materials Based on Functional Group Identity and Substitution Pattern. The Journal of Physical Chemistry Letters 2023, 14 (5) , 1318-1325. https://doi.org/10.1021/acs.jpclett.3c00182
  5. Sambasiva R. Bheemireddy, Zhiguang Li, Jingjing Zhang, Garvit Agarwal, Lily A. Robertson, Ilya A. Shkrob, Rajeev S. Assary, Zhengcheng Zhang, Xiaoliang Wei, Lei Cheng, Lu Zhang. Fluorination Enables Simultaneous Improvements of a Dialkoxybenzene-Based Redoxmer for Nonaqueous Redox Flow Batteries. ACS Applied Materials & Interfaces 2022, 14 (25) , 28834-28841. https://doi.org/10.1021/acsami.2c04926
  6. Donghan Xu, Cuijuan Zhang, Yihan Zhen, Yongdan Li. Liquid Nitrobenzene-Based Anolyte Materials for High-Current and -Energy-Density Nonaqueous Redox Flow Batteries. ACS Applied Materials & Interfaces 2021, 13 (30) , 35579-35584. https://doi.org/10.1021/acsami.1c05564
  7. Jason D. Howard, Rajeev S. Assary, Larry A. Curtiss. Insights into the Interaction of Redox Active Organic Molecules and Solvents with the Pristine and Defective Graphene Surfaces from Density Functional Theory. The Journal of Physical Chemistry C 2020, 124 (5) , 2799-2805. https://doi.org/10.1021/acs.jpcc.9b10403
  8. Elena C. Montoto, Yu Cao, Kenneth Hernández-Burgos, Christo S. Sevov, Miles N. Braten, Brett A. Helms, Jeffrey S. Moore, Joaquín Rodríguez-López. Effect of the Backbone Tether on the Electrochemical Properties of Soluble Cyclopropenium Redox-Active Polymers. Macromolecules 2018, 51 (10) , 3539-3546. https://doi.org/10.1021/acs.macromol.8b00574
  9. Jingjing Zhang, Jinhua Huang, Lily A. Robertson, Rajeev S. Assary, Ilya A. Shkrob, Lu Zhang. Elucidating Factors Controlling Long-Term Stability of Radical Anions for Negative Charge Storage in Nonaqueous Redox Flow Batteries. The Journal of Physical Chemistry C 2018, 122 (15) , 8116-8127. https://doi.org/10.1021/acs.jpcc.8b01434
  10. Jingjing Zhang, R. E. Corman, Jonathon K. Schuh, Randy H. Ewoldt, Ilya A. Shkrob, Lu Zhang. Solution Properties and Practical Limits of Concentrated Electrolytes for Nonaqueous Redox Flow Batteries. The Journal of Physical Chemistry C 2018, 122 (15) , 8159-8172. https://doi.org/10.1021/acs.jpcc.8b02009
  11. Jingjing Zhang, Ilya A. Shkrob, Rajeev S. Assary, Siu on Tung, Benjamin Silcox, Larry A. Curtiss, Levi Thompson, and Lu Zhang . Toward Improved Catholyte Materials for Redox Flow Batteries: What Controls Chemical Stability of Persistent Radical Cations?. The Journal of Physical Chemistry C 2017, 121 (42) , 23347-23358. https://doi.org/10.1021/acs.jpcc.7b08281
  12. Xiaoliang Wei, Wenxiao Pan, Wentao Duan, Aaron Hollas, Zheng Yang, Bin Li, Zimin Nie, Jun Liu, David Reed, Wei Wang, and Vincent Sprenkle . Materials and Systems for Organic Redox Flow Batteries: Status and Challenges. ACS Energy Letters 2017, 2 (9) , 2187-2204. https://doi.org/10.1021/acsenergylett.7b00650
  13. Christo S. Sevov, David P. Hickey, Monique E. Cook, Sophia G. Robinson, Shoshanna Barnett, Shelley D. Minteer, Matthew S. Sigman, and Melanie S. Sanford . Physical Organic Approach to Persistent, Cyclable, Low-Potential Electrolytes for Flow Battery Applications. Journal of the American Chemical Society 2017, 139 (8) , 2924-2927. https://doi.org/10.1021/jacs.7b00147
  14. Kenley M. Pelzer, Lei Cheng, and Larry A. Curtiss . Effects of Functional Groups in Redox-Active Organic Molecules: A High-Throughput Screening Approach. The Journal of Physical Chemistry C 2017, 121 (1) , 237-245. https://doi.org/10.1021/acs.jpcc.6b11473
  15. Robert Löwe, Anna Smith. Contamination in LIB Pouch Cells Promoting Self‐Discharge and Crosstalk. Batteries & Supercaps 2024, 7 (12) https://doi.org/10.1002/batt.202400368
  16. Zhiguang Li, Heonjae Jeong, Xiaoting Fang, Yuyue Zhao, Lily A. Robertson, Jingjing Zhang, Ilya A. Shkrob, Lei Cheng, Xiaoliang Wei, Lu Zhang. Multifaceted effects of ring fusion on the stability of charged dialkoxyarene redoxmers. Journal of Power Sources 2024, 608 , 234689. https://doi.org/10.1016/j.jpowsour.2024.234689
  17. Noufal Merukan Chola, Rajaram K. Nagarale. Quinoxaline derivatives as cathode for aqueous zinc battery. Journal of Solid State Electrochemistry 2024, 28 (2) , 419-431. https://doi.org/10.1007/s10008-023-05689-2
  18. Aleksandr Zaichenko, Andreas J Achazi, Simon Kunz, Hermann A Wegner, Jürgen Janek, Doreen Mollenhauer. Static theoretical investigations of organic redox active materials for redox flow batteries. Progress in Energy 2024, 6 (1) , 012001. https://doi.org/10.1088/2516-1083/ad0913
  19. Xuan Zhou, Abhishek Khetan, Jie Zheng, Mark Huijben, René A. J. Janssen, Süleyman Er. Discovery of lead quinone cathode materials for Li-ion batteries. Digital Discovery 2023, 2 (4) , 1016-1025. https://doi.org/10.1039/D2DD00112H
  20. Andrea Hamza, Flóra B. Németh, Ádám Madarász, Anton Nechaev, Petri M. Pihko, Pekka Peljo, Imre Pápai. N‐Alkylated Pyridoxal Derivatives as Negative Electrolyte Materials for Aqueous Organic Flow Batteries: Computational Screening**. Chemistry – A European Journal 2023, 29 (44) https://doi.org/10.1002/chem.202300996
  21. Lily A. Robertson, Mohammad Afsar Uddin, Ilya A. Shkrob, Jeffrey S. Moore, Lu Zhang. Liquid Redoxmers for Nonaqueous Redox Flow Batteries. ChemSusChem 2023, 16 (14) https://doi.org/10.1002/cssc.202300043
  22. Xuan Zhou, René A. J. Janssen, Süleyman Er. Virtual screening of organic quinones as cathode materials for sodium-ion batteries. Energy Advances 2023, 2 (6) , 820-828. https://doi.org/10.1039/D2YA00282E
  23. Abhishek Khetan. High-Throughput Virtual Screening of Quinones for Aqueous Redox Flow Batteries: Status and Perspectives. Batteries 2023, 9 (1) , 24. https://doi.org/10.3390/batteries9010024
  24. Donghan Xu, Cuijuan Zhang, Yongdan Li. Molecular engineering redox-active organic materials for nonaqueous redox flow battery. Current Opinion in Chemical Engineering 2022, 37 , 100851. https://doi.org/10.1016/j.coche.2022.100851
  25. Maria Giovanna Buonomenna. Membranes for redox flow batteries. 2022, 255-406. https://doi.org/10.1016/B978-0-08-101985-6.00006-9
  26. Jingjing Zhang, Ilya A. Shkrob, Lily A. Robertson, Lu Zhang. Multiple charging and chemical stability of tripodal catholyte redoxmers. Chemical Physics Letters 2022, 787 , 139212. https://doi.org/10.1016/j.cplett.2021.139212
  27. Donghan Xu, Cuijuan Zhang, Yihan Zhen, Yicheng Zhao, Yongdan Li. A high-rate nonaqueous organic redox flow battery. Journal of Power Sources 2021, 495 , 229819. https://doi.org/10.1016/j.jpowsour.2021.229819
  28. Jean-Christophe Daigle, Sylviane Rochon, Yuichiro Asakawa, Benoît Fleutot, Charlotte Mallet, Kamyab Amouzegar, Karim Zaghib. High performance LiMnFePO 4 /Li 4 Ti 5 O 12 full cells by functionalized polymeric additives. Materials Advances 2021, 2 (1) , 253-260. https://doi.org/10.1039/D0MA00679C
  29. Carlos de la Cruz, Antonio Molina, Nagaraj Patil, Edgar Ventosa, Rebeca Marcilla, Andreas Mavrandonakis. New insights into phenazine-based organic redox flow batteries by using high-throughput DFT modelling. Sustainable Energy & Fuels 2020, 4 (11) , 5513-5521. https://doi.org/10.1039/D0SE00687D
  30. Jingjing Zhang, Ilya A. Shkrob, Rajeev S. Assary, Ronald J. Clark, Richard E. Wilson, Sisi Jiang, Quinton J. Meisner, Lei Zhu, Bin Hu, Lu Zhang. An extremely durable redox shuttle additive for overcharge protection of lithium-ion batteries. Materials Today Energy 2019, 13 , 308-311. https://doi.org/10.1016/j.mtener.2019.06.003
  31. Rajeev S. Assary, Larry A. Curtiss. Molecular Level Understanding of the Interactions Between Reaction Intermediates of Li–S Energy Storage Systems and Ether Solvents. 2019, 133-146. https://doi.org/10.1002/9781119297895.ch5
  32. Changkun Zhang, Leyuan Zhang, Yu Ding, Sangshan Peng, Xuelin Guo, Yu Zhao, Gaohong He, Guihua Yu. Progress and prospects of next-generation redox flow batteries. Energy Storage Materials 2018, 15 , 324-350. https://doi.org/10.1016/j.ensm.2018.06.008
  33. Long Huan, Ju Xie, Zhiling Huang, Ming Chen, Guowang Diao, Tongfei Zuo. Computational electrochemistry of Pillar[5]quinone cathode material for lithium-ion batteries. Computational Materials Science 2017, 137 , 233-242. https://doi.org/10.1016/j.commatsci.2017.05.045
  34. Francis Kirby Bokingo Burnea, Hu Shi, Kyoung Chul Ko, Jin Yong Lee. Reduction potential tuning of first row transition metal MIII/MII (M = Cr, Mn, Fe, Co, Ni) hexadentate complexes for viable aqueous redox flow battery catholytes: A DFT study. Electrochimica Acta 2017, 246 , 156-164. https://doi.org/10.1016/j.electacta.2017.05.199
  35. Long Huan, Ju Xie, Ming Chen, Guowang Diao, Rongfang Zhao, Tongfei Zuo. Theoretical investigation of pillar[4]quinone as a cathode active material for lithium-ion batteries. Journal of Molecular Modeling 2017, 23 (4) https://doi.org/10.1007/s00894-017-3282-3
  36. Christo S. Sevov, Sharmila K. Samaroo, Melanie S. Sanford. Cyclopropenium Salts as Cyclable, High‐Potential Catholytes in Nonaqueous Media. Advanced Energy Materials 2017, 7 (5) https://doi.org/10.1002/aenm.201602027

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2016, 120, 27, 14531–14538
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.6b04263
Published June 14, 2016

Copyright © 2016 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

2704

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Schematic of dimethoxybenzene (DMB) derivatives considered in this study.

    Figure 2

    Figure 2. Key properties controlling the stability of redox active species and the quantum chemical descriptors that can assess the properties are schematically shown. Note: G, E, and λ represent Gibbs free energy, redox potential, and reorganization energy, respectively.

    Figure 3

    Figure 3. Schematic of the computation of reorganization energy during oxidation (λox) and reduction (λred) process from the computed energies of a molecule in its neutral and singly positive charge state. Total reorganization energy ((λtotal) is the sum of reorganization energy required during the oxidation (λox) and reduction (λred) process.

    Figure 4

    Figure 4. Comparison of computed oxidation potentials (Eoxn V unit w.r.t. Li/Li+ ref electrode), reorganization energy during oxidation (λox), and root-mean-square deviation (RMSD) between the optimized xyz coordinates of neutral and oxidized states of various DMB derivatives (x-axis). Note: The data associated with the plot is given in Table 1.

    Figure 5

    Figure 5. Schematic of selected reactions (rxn1–4) of 1,4-dimethoxybenzene (DMB) radical cation. Abbreviation: PC, propylene carbonate molecule. All reactions are modeled in aqueous dielectric medium. In rxn1, the cation radical reacts with water to form solvated proton and neutral radical. In rxn4, two cation radicals interact with water to form a dimer and two solvated protons.

    Figure 6

    Figure 6. Computed solution phase enthalpy profile (ΔHsoln) of demethylation of the 25-DMB cation radical by a propylene carbonate molecule. Computed apparent enthalpy of activation (ΔH) of all DMB candidates are given in Table 3.

    Figure 7

    Figure 7. Comparison of the computed free energies changes (ΔG(CH3+)PC) and activation enthalpies (ΔH(CH3+)PC) required for the demethylation of DMB cation radicals by propylene carbonate. The data associated with (ΔG(CH3+)PC) and (ΔH(CH3+)PC) are given in Tables 2 and 3, respectively.

  • References


    This article references 39 other publications.

    1. 1
      Dunn, B.; Kamath, H.; Tarascon, J.-M. Electrical Energy Storage for the Grid: A Battery of Choices Science 2011, 334, 928 935 DOI: 10.1126/science.1212741
    2. 2
      Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi, D.; Lemmon, J. P.; Liu, J. Electrochemical Energy Storage for Green Grid Chem. Rev. 2011, 111, 3577 3613 DOI: 10.1021/cr100290v
    3. 3
      Larcher, D.; Tarascon, J. M. Towards Greener and more Sustainable Batteries for Electrical Energy storage Nat. Chem. 2014, 7, 19 29 DOI: 10.1038/nchem.2085
    4. 4
      Huskinson, B.; Marshak, M. P.; Suh, C.; Er, S.; Gerhardt, M. R.; Galvin, C. J.; Chen, X.; Aspuru-Guzik, A.; Gordon, R. G.; Aziz, M. J. A Metal-free Organic-Inorganic Aqueous Flow Battery Nature 2014, 505, 195 198 DOI: 10.1038/nature12909
    5. 5
      Lin, K.; Chen, Q.; Gerhardt, M. R.; Tong, L.; Kim, S. B.; Eisenach, L.; Valle, A. W.; Hardee, D.; Gordon, R. G.; Aziz, M. J.; Marshak, M. P. Alkaline Quinone Flow Battery Science 2015, 349, 1529 1532 DOI: 10.1126/science.aab3033
    6. 6
      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. 2014, 161, A1371 A1380 DOI: 10.1149/2.1001409jes
    7. 7
      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. 2016, 6, 1501449 1501456 DOI: 10.1002/aenm.201501449
    8. 8
      Wei, X.; Cosimbescu, L.; Xu, W.; Hu, J. Z.; Vijayakumar, M.; Feng, J.; Hu, M. Y.; Deng, X.; Xiao, J.; Liu, J.; Sprenkle, V.; Wang, W. Towards High-Performance Nonaqueous Redox Flow Electrolyte Via Ionic Modification of Active Species Adv. Energy Mater. 2015, 5, 1400678 1400685 DOI: 10.1002/aenm.201400678
    9. 9
      Brushett, F. R.; Vaughey, J. T.; Jansen, A. N. An All-Organic Non-aqueous Lithium-Ion Redox Flow Battery Adv. Energy Mater. 2012, 2, 1390 1396 DOI: 10.1002/aenm.201200322
    10. 10
      Gong, K.; Fang, Q.; Gu, S.; Li, S. F. Y.; Yan, Y. Nonaqueous Redox-flow Batteries: Organic Solvents, Supporting electrolytes, and Redox pairs Energy Environ. Sci. 2015, 8, 3515 3530 DOI: 10.1039/C5EE02341F
    11. 11
      Sevov, C. S.; Brooner, R. E. M.; Chénard, E.; Assary, R. S.; Moore, J. S.; Rodríguez-López, J.; Sanford, M. S. Evolutionary Design of Low Molecular Weight Organic Anolyte Materials for Applications in Nonaqueous Redox Flow Batteries J. Am. Chem. Soc. 2015, 137, 14465 14472 DOI: 10.1021/jacs.5b09572
    12. 12
      Buhrmester, C.; Chen, J.; Moshurchak, L.; Jiang, J.; Wang, R. L.; Dahn, J. R. Studies of Aromatic Redox Shuttle Additives for LiFePO4-Based Li-Ion Cells J. Electrochem. Soc. 2005, 152, A2390 A2399 DOI: 10.1149/1.2098265
    13. 13
      Chen, J.; Buhrmester, C.; Dahn, J. R. Chemical Overcharge and Overdischarge Protection for Lithium-Ion Batteries Electrochem. Solid-State Lett. 2005, 8, A59 A62 DOI: 10.1149/1.1836119
    14. 14
      Zhang, L.; Zhang, Z.; Redfern, P. C.; Curtiss, L. A.; Amine, K. Molecular Engineering towards Safer Lithium-ion Batteries: A highly Stable and Compatible Redox shuttle for Overcharge protection Energy Environ. Sci. 2012, 5, 8204 8207 DOI: 10.1039/c2ee21977h
    15. 15
      Zhang, L.; Zhang, Z.; Wu, H.; Amine, K. Novel redox shuttle Additive for High-Voltage Cathode Materials Energy Environ. Sci. 2011, 4, 2858 2862 DOI: 10.1039/c0ee00733a
    16. 16
      Zhang, Z.; Zhang, L.; Schlueter, J. A.; Redfern, P. C.; Curtiss, L.; Amine, K. Understanding the Redox shuttle stability of 3,5-di-tert-butyl-1,2-dimethoxybenzene for Overcharge Protection of Lithium-ion Batteries J. Power Sources 2010, 195, 4957 4962 DOI: 10.1016/j.jpowsour.2010.02.075
    17. 17
      Speelman, A. L.; Gillmore, J. G. Efficient Computational Methods for Accurately Predicting Reduction Potentials of Organic Molecules J. Phys. Chem. A 2008, 112, 5684 5690 DOI: 10.1021/jp800782e
    18. 18
      Er, S.; Suh, C.; Marshak, M. P.; Aspuru-Guzik, A. Computational Design of Molecules for an all-Quinone Redox Flow Battery Chem. Sci. 2015, 6, 885 893 DOI: 10.1039/C4SC03030C
    19. 19
      Assary, R. S.; Brushett, F. R.; Curtiss, L. A. Reduction Potential Predictions of some Aromatic Nitrogen-Containing Molecules RSC Adv. 2014, 4, 57442 57451 DOI: 10.1039/C4RA08563A
    20. 20
      Bachman, J. E.; Curtiss, L. A.; Assary, R. S. Investigation of the Redox Chemistry of Anthraquinone Derivatives Using Density Functional Theory J. Phys. Chem. A 2014, 118, 8852 8860 DOI: 10.1021/jp5060777
    21. 21
      Hernández-Burgos, K.; Burkhardt, S. E.; Rodríguez-Calero, G. G.; Hennig, R. G.; Abruña, H. D. Theoretical Studies of Carbonyl-Based Organic Molecules for Energy Storage Applications: The Heteroatom and Substituent Effect J. Phys. Chem. C 2014, 118, 6046 6051 DOI: 10.1021/jp4117613
    22. 22
      Pineda Flores, S. D.; Martin-Noble, G. C.; Phillips, R. L.; Schrier, J. Bio-Inspired Electroactive Organic Molecules for Aqueous Redox Flow Batteries. 1. Thiophenoquinones J. Phys. Chem. C 2015, 119, 21800 21809 DOI: 10.1021/acs.jpcc.5b05346
    23. 23
      Karlsson, C.; Jämstorp, E.; Strømme, M.; Sjödin, M. Computational electrochemistry study of 16 isoindole-4, 7-diones as candidates for organic cathode materials J. Phys. Chem. C 2012, 116, 3793 3801 DOI: 10.1021/jp211851f
    24. 24
      Coote, M. L.; Lin, C. Y.; Beckwith, A. L. J.; Zavitsas, A. A. A comparison of methods for measuring relative radical stabilities of carbon-centred radicals Phys. Chem. Chem. Phys. 2010, 12, 9597 9610 DOI: 10.1039/c003880f
    25. 25
      Cheng, L.; Assary, R. S.; Qu, X.; Jain, A.; Ong, S. P.; Rajput, N. N.; Persson, K.; Curtiss, L. A. Accelerating Electrolyte Discovery for Energy Storage with High-Throughput Screening J. Phys. Chem. Lett. 2015, 6, 283 291 DOI: 10.1021/jz502319n
    26. 26
      Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Gaussian 09; Gaussian, Inc.: Wallingford, CT, U.S.A., 2009.
    27. 27
      Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Aqueous Solvation Free Energies of Ions and Ion-Water Clusters Based on an Accurate Value for the Absolute Aqueous Solvation Free Energy of the Proton J. Phys. Chem. B 2006, 110, 16066 16081 DOI: 10.1021/jp063552y
    28. 28
      Bhattacharyya, S.; Stankovich, M. T.; Truhlar, D. G.; Gao, J. Combined Quantum Mechanical and Molecular Mechanical Simulations of One- and Two-Electron Reduction Potentials of Flavin Cofactor in Water, Medium-Chain Acyl-CoA Dehydrogenase, and Cholesterol Oxidase J. Phys. Chem. A 2007, 111, 5729 5742 DOI: 10.1021/jp071526+
    29. 29
      Guerard, J. J.; Arey, J. S. Critical Evaluation of Implicit Solvent Models for Predicting Aqueous Oxidation Potentials of Neutral Organic Compounds J. Chem. Theory Comput. 2013, 9, 5046 5058 DOI: 10.1021/ct4004433
    30. 30
      Moens, J.; Geerlings, P.; Roos, G. A Conceptual DFT Approach for the Evaluation and Interpretation of Redox Potentials Chem. - Eur. J. 2007, 13, 8174 8184 DOI: 10.1002/chem.200601896
    31. 31
      Borodin, O.; Behl, W.; Jow, T. R. Oxidative Stability and Initial Decomposition Reactions of Carbonate, Sulfone, and Alkyl Phosphate-Based Electrolytes J. Phys. Chem. C 2013, 117, 8661 8682 DOI: 10.1021/jp400527c
    32. 32
      Vollmer, J. M.; Curtiss, L. A.; Vissers, D. R.; Amine, K. Reduction Mechanisms of Ethylene, Propylene, and Vinylethylene Carbonates: A Quantum Chemical Study J. Electrochem. Soc. 2004, 151, A178 A183 DOI: 10.1149/1.1633765
    33. 33
      Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Single-Ion Solvation Free Energies and the Normal Hydrogen Electrode Potential in Methanol, Acetonitrile, and Dimethyl Sulfoxide J. Phys. Chem. B 2007, 111, 408 422 DOI: 10.1021/jp065403l
    34. 34
      Ho, J.; Coote, M. L. First-principles prediction of acidities in the gas and solution phase Wiley Interdisc. Rev.: Comput. Mol. Sci. 2011, 1, 649 660 DOI: 10.1002/wcms.43
    35. 35
      Assary, R. S.; Curtiss, L. A.; Moore, J. S. Toward a Molecular Understanding of Energetics in Li–S Batteries Using Nonaqueous Electrolytes: A High-Level Quantum Chemical Study J. Phys. Chem. C 2014, 118, 11545 11558 DOI: 10.1021/jp5015466
    36. 36
      Kabsch, W. A solution for the Best Rotation to Relate Two Sets of Vectors Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1976, 32, 922 923 DOI: 10.1107/S0567739476001873
    37. 37
      Schmittel, M.; Burghart, A. Understanding Reactivity Patterns of Radical Cations Angew. Chem., Int. Ed. Engl. 1997, 36, 2550 2589 DOI: 10.1002/anie.199725501
    38. 38
      Kelly, C. P.; Cramer, C. J.; Truhlar, D. G. Single-Ion Solvation Free Energies and the Normal Hydrogen Electrode Potential in Methanol, Acetonitrile, and Dimethyl Sulfoxide J. Phys. Chem. B 2007, 111, 408 422 DOI: 10.1021/jp065403l
    39. 39
      Assary, R. S.; Kim, T.; Low, J. J.; Greeley, J.; Curtiss, L. A. Glucose and Fructose to Platform Chemicals: Understanding the Thermodynamic Landscapes of Acid-Catalysed Reactions Using High-Level Ab Initio Methods Phys. Chem. Chem. Phys. 2012, 14, 16603 16611 DOI: 10.1039/c2cp41842h
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.6b04263.

    • Optimized structures of DMB derivatives are shown in Figure S1. Figure 2 suggests most important structural parameters and these parameters are tabulated in Table S1. Figure S3 presents the computed deprotonation energies (rxn 1 of Figure 5) and dimerization energies (rxn 4 of the Figure 5) of radical cations at the B3LYP/6-31G(2df,p) level of theory, and Table S2 shows computed enthalpies and free energies of activation required for the demethylation reactions in water and diethyl ether dielectric media (PDF).


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.