Biopolymeric Carbon as Electrode and ILs as Electrolytes: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Electrochemical capacitors (also called supercapacitors) are an important segment of the clean energy portfolio due to their high-power density and long cycle life. The energy storage mechanism in supercapacitors is based on the electric charge buildup by a charge–discharge process (electrosorption) at the electrode–electrolyte interface and/or redox reactions on the electroactive surface.

  • biopolymers
  • EDLCs
  • flexible electrodes
  • gel polymer electrolytes

1. Historic Perspective of IL Electrolytes for Electric Double Layer Capacitors (EDLCs)

1.1. Performance Requirements for EDLCs: The Emergence of ILs as Electrolytes

Electrochemical capacitors (also called supercapacitors) are an important segment of the clean energy portfolio due to their high-power density and long cycle life. The energy storage mechanism in supercapacitors is based on the electric charge buildup by a charge–discharge process (electrosorption) at the electrode–electrolyte interface and/or redox reactions on the electroactive surface. The energy storage in supercapacitors occurs via two complementary types of contributions, i.e., contributions from electric-double layer capacitance and contributions from pseudocapacitive faradaic processes. The distribution of these two types of capacitances depends on both the material and structure of the electrodes. Hybrid pseudocapacitors utilizing a faradaic process at one electrode and a purely capacitive component at the other are also known. In addition to the electrode material, the electrolyte is an important part of the supercapacitor and affects both the energy and power performance of the device.
Efforts to miniaturize supercapacitors have mainly focused on purely electric double-layer capacitors (EDLCs)—high power density devices, using electrodes with high surface areas based on traditional porous carbon, graphene [1], reduced graphene oxide/carbon nanotubes [2], onion-like carbon [3], etc. A few miniaturized systems have used RuO2 [4] or conjugated polymers [5][6][7] to add a pseudocapacitive component to the energy-storage system, although these polymers are not optimal for high-energy density systems.
Miniaturized supercapacitors present two fundamental challenges: 1. While the power density is high, the energy density is low, and 2. The current is delivered at a continuously decreasing voltage. The power (P) of EDLCs is proportional to the operating voltage (V) and inversely proportional to the internal resistance (ESR). While ESR depends on the design of the ultracapacitor and the types of electrodes, V depends on the stability of the electrolyte at the applied potential. Thus, a proper choice of electrolytes is the most effective way to increase both the energy and the power of EDLCs. However, studies have shown that it is not possible to fabricate EDLCs with voltages higher than 3 V using conventional organic electrolytes. In addition, conventional electrolytes suffer from operational safety problems related to volatility and flammability. Hence, in the last decade, much effort has been dedicated to the development of safer electrolytes with larger electrochemical windows (>3 V).
Electrolytes based on ionic liquids (ILs) are the most promising [8]. In fact, the emergence of the IL field is, in part, a legacy of the United States Air Force efforts that conducted electrochemical studies using room-temperature molten salts [9][10]. ILs offer unique material properties, including thermal and chemical stability, a broad electrochemical window, high ionic conductivity, and non-flammability. All these properties can be easily modified by tuning the component ions of IL electrolytes [11][12][13][14]. Redox-stable ILs with large electrochemical windows have led to a resurgence in interest in electrochemical applications with promising technological applications [15][16][17].
As electrolytes, ILs have preferential characteristics such as high conductivity (mostly up to 2 S m−1, with exceptionally high conductivities of 10 S m−1 being reported for imidazolium-based ILs [18]). A comprehensive analysis and correlation of IL structure vs. conductivity data for energy applications indicated that most of the evaluated ILs presented conductivities within the range of 7 × 10−3 to 7 S m−1, whereas a conductivity of the ‘classic’ battery electrolyte Li[PF6] solution ranges from 0.7 to 1.5 S m−1 [19]. The second attractive property of the ILs is their wide electrochemical stability window (up to 6.3 V [20][21][22][23][24][25]), with no degradation of electrochemical performance compared with the traditional electrolytes. Another important bulk characteristic of ILs is their static dielectric constant (εS), a characteristic that describes ILs’ solvation capability, which depends on the cation–anion combination and has been determined in aprotic ILs with [NTf2]-anion to be in the range of 12–15.8, and in ILs with [C2mim]-cation 11.7–35.0 [20]. Static dielectric constants of some protic ILs have εS values up to 85.6 [20]. On the other hand, the (often) high viscosity of the ILs might hinder charge transport and lower transference number. In addition, the capacitance/behavior of ILs at electrodes (types of electrodes and the layering behavior of the IL electrolyte) must be taken into consideration for supercapacitor applications.

2. Biopolymeric Carbon as Electrode and ILs as Electrolytes

In addition to using biopolymers as main components in solid electrolytes, biopolymers can be used to create electrodes that would be advantageous as a freestanding and binder-free type of electrode for flexible supercapacitors. 
Because a conventional binder uses synthetic polymers [26][27], it can be replaced with a biopolymer-based binder consisting of cellulose-based materials [28], chitin/chitosan [29][30], lignin [31], etc., able to hold other materials together and exhibiting good bonding properties with the current collector. ILs can simultaneously dissolve biopolymers and stabilize nanoparticles. For instance, graphene is rich in π-electrons, and it has been shown that strong IL cation–π interactions exist between carbon nanomaterial and IL with aromatic cations such as imidazolium. It was recently shown with computational studies that the interaction of the ILs with the graphitic surfaces is considerably influenced by the charge transfer between the component ions [32]. Multiple approaches that allow the formation of composites of various biopolymers with nanomaterials to produce films and membranes have been reported, utilizing activated carbon [33], graphene [29][34], graphene oxide [35], polypyrrole (PPy) [36], multiwalled carbon nanotubes (MWCNTs) [37], etc. Still, only a few reports can be found using biopolymer-based electrodes in combination with ILs as electrolytes. For example, cellulose was dissolved in the IL [C4mim]Cl and combined with PPy and trihexyl(tetradecyl)phosphonium bis[(trifluoromethyl)sulfonyl] amide ([P66614][NTf2]) as the IL plasticizer to form PPy–Cellulose–IL composite films via light-induced polymerization [36].
Porous carbon is one of the best electrode materials to date and found to be widely used to synthesize an efficient supercapacitor device. Porous carbon has outstanding properties such as higher surface area, larger pore volume, great electrical conductivity, and efficient chemical stability. Also, it is environmentally friendly and can be made from any organic matter on Earth. A simple, costless, and nonpolluting strategy was proposed by using the decomposable and water-removable NaNO3 salt crystals as both the template and pore engineer in the gelatin biopolymer aerogel to fabricate a cross-coupled macro-mesoporous carbon material, with high surface area (approaching 3000 m2 g−1) [38]. When combined with [C2mim][BF4] as the electrolyte, a high energy density of 92 Wh kg−1 was obtained at 1 kW kg−1, and remained at 39 Wh kg−1 even when the power density increased up to 200 kW kg−1, outperforming nearly all hitherto reported porous carbon at high current density. The calculated specific capacitances of the material were 166, 152, 142, 135, 127, 118, 111, 98, 84, and 70 F g−1 under the current densities of 0.5, 1, 2, 5, 10, 20, 30, 50, 75, and 100 A g−1, respectively. Alternatively, porous carbon was prepared from cornstarch biopolymer by a simple carbonization process, initially at 800 °C [39]. The resulting material was coated with a polymer electrolyte film of PVDF–HFP, doped with 300 wt% of 1-ethyl-3-methylimidazolium tricyanomethanide ([C2mim][TCM]) IL as a separator. A supercapacitor device was then fabricated at a laboratory scale with the prepared porous carbon electrodes sandwiched around the electrolyte film, which yielded a specific capacitance of 188.4 F gm−1 at 10 mHz, confirmed from the electrochemical low-frequency impedance spectroscopy plot. Cyclic voltammetry results showed a high specific capacitance of 184.8 F gm−1 at 5 mV s−1.
Another approach is the thermal processing of different lignin grades into high-performance carbon materials [40][41][42]. Activated porous lignin-based carbons with specific high surface areas of more than 1800 m2 g−1 were recently synthesized by employing a simple two-step process, which consisted of a high-temperature thermal treatment of a lignin/KOH composite under an inert gas atmosphere without any addition of templating agents; the synthesis was followed by a washing step to remove byproducts of the activation procedure [43]. The electrochemical performance of the resulting carbons indicated that these could be used as an active material in double-layer capacitors, using the IL [C2mim][BF4] as the electrolyte to enhance storage ability. A capacitance of 231 F g−1 at 1 A g−1 and 203 F g−1 when the current was increased 10-fold to 10 A g−1 was achieved for carbon with a specific surface area of more than 1800 m2 g−1. One of the most crucial factors determining the electrochemical response of the active materials was found to be the strong surface functionalization by oxygen-containing groups. However, over the course of 10,000 charging−discharging cycles, a decay in capacitance of about 50% was observed, which might be due to the large voltage window and the surface functionalization.
A novel material platform based on choline IL-functionalized biopolymers, which can form a hydrogel electrolyte when exposed to visible light, was also developed. The polymer electrolyte entailed mixing a methacrylate polymer (gelatin methacryloyl, GelMA, or polyethylene glycol diacrylate, PEGDA) and choline acrylate ([Cho][Acrylate]) to make [Cho][Acrylate]–GelMA (BG) and [Cho][Acrylate]–PEGDA (BP) hydrogels, respectively [44]. Lithium phenyl-2,4,6-trimethylbenzoylphosphinate (LAP) was added as a photoinitiator, and visible light 405 nm was used for 60 s. Graphene hydrogel was prepared by reducing graphene oxide with ascorbic acid for use as electrodes. Fine-structure, interdigitated, biocompatible, and implantable soft micro-supercapacitors (MSC) were created by 3D in situ bioprinting of these polymer electrolytes in combination with rheologically-optimized graphene hydrogel–laponite (GH–L) blend as electrode material. The hydrogel electrolyte had a specific capacitance of ~200 F g−1, while the MSC had a specific capacitance of ~16 μF g−1 at a current density of 1 A g−1, volumetric capacitance of ~44 μF cm−3, cyclic stability up to 10,000 cycles, energy densities nearly as high as implantable batteries, and the power density level of implantable supercapacitors. This novel material platform enables in situ 3D printing of flexible bioelectronics structures with an integrated life-long power source.
In all aforementioned cases, it is important to note the electrostatic adsorption of ionic species at the interface of electrode and solution in EDLCs. Electrolytes are typically confined in the pores of electrodes (e.g., carbon), where the confinement region is often smaller than 1 nm, so that resulting pore sizes are only one or two times the diameters of bare ions or solvent molecules [45]. (The ionic size of the common cation and anion of IL electrolytes—[C2mim]+, [C4mim]+, [C1C4Pyr], [N2222]+ paired with fluorinated [NTf2], [DCA], [OTf], and [PF6]—can be found in reference [46].) Hence, the in-pore composition of ions is anticipated to differ from the bulk IL: the ions closest to the electrodes can adsorb and order at the electrode surface [47]. Properties of the system (e.g., charging rates) could be affected by pore width due to variations in initial ion populations, and could be adjusted by modification of the pore diameter of the electrodes. In electrodes with a sub-nanometer size, the pores (so-called ionophobic pores) that contain a small number of ions at zero charge bring about enhanced power and energy density in supercapacitors [48], due to an increase in ion solvation. Recent molecular dynamics simulations on two IL electrolytes, [C2mim][BF4] and [C4mim][NTf2] (1.5 M in acetonitrile), demonstrated that, under confinement between graphene sheets forming slit pores of various widths, in-pore mole fraction of the IL varies from nearly 0 (pore width 0.76 nm) to 0.3 (pore width 0.84 nm), then again decreases to almost 0 (pore width ~1 nm), and then rises again, to accommodate average IL mole fraction of 0.12. Interestingly, the IL composition was important: for [C2mim][BF4]/acetonitrile solution (IL cation width 0.76 nm, anion size 0.45 nm [46]), ions were nearly fully excluded from pores with widths near 1 nm, while for [C4mim][NTf2] (IL cation width 0.90 nm, anion size 0.79 nm [46]), only a slight depletion of ions was observed. In addition, the IL size asymmetry should be considered: the spontaneous structure formation at the interface of electrode/electrolyte is affected by ionic size asymmetry, which plays a significant role in charge screening and, hence, ionic density [49].

This entry is adapted from the peer-reviewed paper 10.3390/ijms24097866

References

  1. Liu, C.; Yu, Z.; Neff, D.; Zhamu, A.; Jang, B.Z. Graphene-based supercapacitor with an ultrahigh energy density. Nano Lett. 2010, 10, 4863–4868.
  2. Beidaghi, M.; Wang, C. Micro-supercapacitors based on interdigital electrodes of reduced graphene oxide and carbon nanotube composites with ultrahigh power handling performance. Adv. Funct. Mater. 2012, 22, 4501–4510.
  3. Pech, D.; Brunet, M.; Durou, H.; Huang, P.; Mochalin, V.; Gogotsi, Y.; Taberna, P.-L.; Simon, P. Ultrahigh-power micrometre-sized supercapacitors based on onion-like carbon. Nat. Nanotechnol. 2010, 5, 651–654.
  4. Kim, H.-K.; Cho, S.-H.; Ok, Y.-W.; Seong, T.-Y.; Yoon, Y.S. All solid-state rechargeable thin-film microsupercapacitor fabricated with tungsten co-sputtered ruthenium oxide electrodes. J. Vac. Sci. Technol. B 2003, 21, 949–952.
  5. Wang, K.; Zou, W.; Quan, B.; Yu, A.; Wu, H.; Jiang, P.; Wei, Z. An all-solid-state flexible micro-supercapacitor on a chip. Adv. Energy Mater. 2011, 1, 1068–1072.
  6. Beidaghi, M.; Wang, C. Micro-supercapacitors Based on three dimensional interdigital polypyrrole/C-MEMS electrodes. Electrochim. Acta 2011, 56, 9508–9514.
  7. Khomenko, V.; Frackowiak, E.; Béguin, F. Determination of the specific capacitance of conducting polymer/nanotubes composite electrodes using different cell configurations. Electrochim. Acta 2005, 50, 2499–2506.
  8. Brandt, A.; Pohlmann, S.; Varzi, A.; Balducci, A.; Passerini, S. Ionic Liquids in Supercapacitors. MRS Bull. 2013, 38, 554–559.
  9. Wilkes, J.S.; Levisky, J.A.; Wilson, R.A.; Hussey, C.L. Dialkylimidazolium chloroaluminate melts: A new class of room-temperature ionic liquids for electrochemistry, spectroscopy and synthesis. Inorg. Chem. 1982, 21, 1263–1264.
  10. Osteryoung, R.A.; Osteryoung, J.G. Studies in ambient temperature ionic liquids. In AFOSR Report 1987; AFOSR-TR-87-0879; Order No. AD-A182723; AFOSR: Arlington, VA, USA, 1987.
  11. Zhang, L.; Gao, H.; Jin, G.; Liu, S.; Wu, J.; Wu, H.; Yang, Y.; Wang, Q.; Wang, S. Cellulose-Based Electrolytes for Advanced Lithium-Ion Batteries: Recent Advances and Future Perspectives. ChemNanoMat 2022, 8, e202200142.
  12. Karuppasamy, K.; Theerthagiri, J.; Vikraman, D.; Yim, C.J.; Hussain, S.; Sharma, R.; Maiyalagan, T.; Qin, J.; Kim, H.S. Ionic liquid-based electrolytes for energy storage devices: A brief review on their limits and applications. Polymers 2020, 12, 918.
  13. MacFarlane, D.R.; Forsyth, M.; Howlett, P.C.; Kar, M.; Passerini, S.; Pringle, J.M.; Ohno, H.; Watanabe, M.; Yan, F.; Zheng, W.; et al. Ionic liquids and their solid-state analogues as materials for energy generation and storage. Nat. Rev. Mater. 2016, 1, 15005.
  14. Varela, J.C.; Sankar, K.; Hino, A.; Lin, X.; Chang, W.-s.; Coker, D.; Grinstaff, M. Piperidinium ionic liquids as electrolyte solvents for sustained high temperature supercapacitor operation. Chem. Commun. 2018, 54, 5590–5593.
  15. Pitawala, J.; Navarra, M.A.; Scrosati, B.; Jacobsson, P.; Matic, A. Structure and properties of Li-ion conducting polymer gel electrolytes based on ionic liquids of the pyrrolidinium cation and the bis (trifluoromethanesulfonyl) imide anion. J. Power Sources 2014, 245, 830–835.
  16. Srour, H.; Rouault, H.; Santini, C. Imidazolium based ionic liquid electrolytes for Li-ion secondary batteries based on graphite and LiFePO4. J. Electrochem. Soc. 2012, 160, A66–A69.
  17. Mousavi, M.P.S.; Wilson, B.E.; Kashefolgheta, S.; Anderson, E.L.; He, S.; Bühlmann, P.; Stein, A. Ionic liquids as electrolytes for electrochemical double-layer capacitors: Structures that optimize specific energy. ACS Appl. Mater. Interfaces 2016, 8, 3396–3406.
  18. Galiński, M.; Lewandowski, A.; Stępniak, I. Ionic liquids as electrolytes. Electrochim. Acta 2006, 51, 5567–5580.
  19. Nancarrow, P.; Al-Othman, A.; Mital, D.K.; Dopking, S. Comprehensive analysis and correlation of ionic liquid conductivity data for energy applications. Energy 2021, 220, 119761.
  20. Weingärtner, H. The static dielectric permittivity of ionic liquids. J. Molec. Liquid 2014, 192, 185–190.
  21. Kalb, R.S. Chapter 11: Toward Industrialization of Ionic Liquids. In Commercial Applications of Ionic Liquids; Shiflett, M., Ed.; Springer Nature: Cham, Switzerland, 2020.
  22. Matsumoto, H.; Sakaebe, H.; Tatsumi, K.; Kikuta, M.; Ishiko, E.; Kono, M. Fast cycling of Li/LiCoO2 cell with low-viscosity ionic liquids based on bis(fluorosulfonyl)imide −. J. Power Sources 2006, 160, 1308–1313.
  23. Zhou, Q.; Henderson, W.A.; Appetecchi, G.B.; Montanino, M.; Passerini, S. Physical and electrochemical properties of n-alkyl-n-methylpyrrolidinium bis(fluorosulfonyl)imide ionic liquids: PY13FSI and PY14FSI. J. Phys. Chem. B 2008, 112, 13577–13580.
  24. Liu, K.; Zhou, Y.-X.; Han, H.-B.; Zhou, S.-S.; Feng, W.-F.; Nie, J.; Li, H.; Huang, X.-J.; Armand, M.; Zhou, Z.-B. Ionic liquids based on (fluorosulfonyl)(pentafluoroethanesulfonyl)imide with various oniums. Electrochim. Acta 2010, 55, 7145–7151.
  25. MacFarlane, D.R.; Meakin, P.; Sun, J.; Amini, N.; Forsyth, M. Pyrrolidinium imides: A new family of molten salts and conductive plastic crystal phases. J. Phys. Chem. B 1999, 103, 4164–4170.
  26. Abdah, M.A.A.M.; Zubair, N.A.; Azman, N.H.N.; Sulaiman, Y. Fabrication of PEDOT coated PVA-GO nanofiber for supercapacitor. Mater. Chem. Phys. 2017, 192, 161.
  27. Dong, J.; Wang, Z.; Kang, X. The synthesis of graphene/PVDF composite binder and its application in high performance MnO2 supercapacitors. Colloids Surf. A Physicochem. Eng. Asp. 2016, 489, 282–288.
  28. Andres, B.; Dahlström, C.; Blomquist, N.; Norgren, M.; Olin, H. Cellulose binders for electric double-layer capacitor electrodes: The influence of cellulose quality on electrical properties. Mater. Des. 2018, 141, 342–349.
  29. Rogers, R.D.; Zavgorodnya, O.; Shamshina, J.L.; Gurau, G. Graphene-Biopolymer Composite Materials and Methods of Making Thereof. U.S. Patent 15/936,056, 2018.
  30. Choudhury, N.A.; Northrop, P.W.; Crothers, A.C.; Jain, S.; Subramanian, V.R. Chitosan hydrogel-based electrode binder and electrolyte membrane for EDLCs: Experimental studies and model validation. J. Appl. Electrochem. 2012, 42, 935–943.
  31. Zhao, Z.; Cannon, F.S.; Nieto-Delgado, C.; Pena, L. Lignin/collagen hybrid biomaterials as binder substitute for specialty graphites and electrodes. Carbon 2016, 108, 303–317.
  32. Ghatee, M.H.; Moosavi, F. Physisorption of hydrophobic and hydrophilic 1-alkyl-3-methylimidazolium ionic liquids on the graphenes. J. Phys. Chem. C 2011, 115, 5626–5636.
  33. Murashko, K.; Nevstrueva, D.; Pihlajamäki, A.; Koiranen, T.; Pyrhönen, J. Cellulose and activated carbon based flexible electrical double-layer capacitor electrode: Preparation and characterization. Energy 2017, 119, 435–441.
  34. Ye, W.; Li, X.; Zhu, H.; Wang, X.; Wang, S.; Wang, H.; Sun, R. Green fabrication of cellulose/graphene composite in ionic liquid and its electrochemical and photothermal properties. Chem. Eng. J. 2016, 299, 45–55.
  35. King, C.; Easton, M.E.; Rogers, R.D. Chitin for the replacement of fluoropolymers in the assembly of electrochemical devices. ChemRxiv Camb. Camb. Open Engag. 2018.
  36. Liu, Y.; Wang, Y.; Nie, Y.; Wang, C.; Ji, X.; Zhou, L.; Pan, F.; Zhang, S. Preparation of MWCNTs-graphene-cellulose fiber with ionic liquids. ACS Sustain. Chem. Eng. 2019, 7, 20013–22002.
  37. Lorenzo, M.; Srinivasan, G. Durable flexible supercapacitors utilizing the multifunctional role of ionic liquids. Energy Technol. 2018, 6, 196–204.
  38. Li, J.; Wang, N.; Tian, J.; Qian, W.; Chu, W. Cross-coupled macro-mesoporous carbon network toward record high energy-power density supercapacitor at 4 V. Adv. Funct. Mater. 2018, 28, 1806153.
  39. Nath, G.; Singh, P.K.; Dhapola, P.S.; Dohare, S.; Noor, I.M.; Sharma, T.; Singh, A. Fabrication of cornstarch biopolymer-derived nano porous carbon as electrode material for supercapacitor application. Biomass Convers. Biorefinery 2022, 1–8.
  40. Suhas; Carrott, P.J.M.; Ribeiro Carrott, M.M.L. Lignin-from natural adsorbent to activated carbon: A review. Bioresour. Technol. 2007, 98, 2301–2312.
  41. Babeł, K.; Jurewicz, K. KOH activated lignin based nanostructured carbon exhibiting high hydrogen electrosorption. Carbon 2008, 46, 1948–1956.
  42. Kijima, M.; Hirukawa, T.; Hanawa, F.; Hata, T. Thermal conversion of alkaline lignin and its structured derivatives to porous carbonized materials. Bioresour. Technol. 2011, 102, 6279–6285.
  43. Klose, M.; Reinhold, R.; Logsch, F.; Wolke, F.; Linnemann, J.; Stoeck, U.; Oswald, S.; Uhlemann, M.; Balach, J.; Markowski, J.; et al. Softwood lignin as a sustainable feedstock for porous carbons as active material for supercapacitors using an ionic liquid electrolyte. ACS Sustain. Chem. Eng. 2017, 5, 4094–4102.
  44. Krishnadoss, V.; Kanjilal, B.; Hesketh, A.; Miller, C.; Mugweru, A.; Akbard, M.; Khademhosseini, A.; Leijten, J.; Noshadi, I. In situ 3D printing of implantable energy storage devices. Chem. Eng. J. 2021, 409, 128213.
  45. Fang, A.; Smolyanitsky, A. Large variations in the composition of ionic liquid-solvent mixtures in nanoscale confinement. ACS Appl. Mater. Interfaces 2019, 11, 27243–27250.
  46. Sun, L.; Zhuo, K.; Chen, Y.; Du, Q.; Zhang, S.; Wang, J. Ionic liquid-based redox active electrolytes for supercapacitors. Adv. Funct. Mater. 2022, 32, 2203611.
  47. Kislenko, S.A.; Samoylov, I.S.; Amirov, R.H. Molecular dynamics simulation of the electrochemical interface between a graphite surface and the ionic liquid . Phys. Chem. Chem. Phys. 2009, 11, 5584–5590.
  48. Kondrat, S.; Wu, P.; Qiao, R.; Kornyshev, A.A. Accelerating charging dynamics in subnanometre pores. Nat. Mater 2014, 13, 387.
  49. de Souza, J.P.; Pivnic, K.; Bazant, M.Z.; Urbakh, M.; Kornyshev, A.A. Structural forces in ionic liquids: The role of ionic size asymmetry. J. Phys. Chem. B 2022, 126, 1242–1253.
More
This entry is offline, you can click here to edit this entry!
Video Production Service