Solid state ionics

Last updated
Power density vs. energy density for different classes of solid state ionics systems used for energy storage and conversion Power-energy for energy storage systems.png
Power density vs. energy density for different classes of solid state ionics systems used for energy storage and conversion

Solid-state ionics is the study of ionic-electronic mixed conductor and fully ionic conductors (solid electrolytes) and their uses. Some materials that fall into this category include inorganic crystalline and polycrystalline solids, ceramics, glasses, polymers, and composites. Solid-state ionic devices, such as solid oxide fuel cells, can be much more reliable and long-lasting, especially under harsh conditions, than comparable devices with fluid electrolytes. [1]

Contents

The field of solid-state ionics was first developed in Europe, starting with the work of Michael Faraday on solid electrolytes Ag2S and PbF2 in 1834. Fundamental contributions were later made by Walther Nernst, who derived the Nernst equation and detected ionic conduction in heterovalently doped zirconia, which he applied in his Nernst lamp. Another major step forward was the characterization of silver iodide in 1914. Around 1930, the concept of point defects was established by Yakov Frenkel, Walter Schottky and Carl Wagner, including the development of point-defect thermodynamics by Schottky and Wagner; this helped explain ionic and electronic transport in ionic crystals, ion-conducting glasses, polymer electrolytes and nanocomposites. In the late 20th and early 21st centuries, solid-state ionics focused on the synthesis and characterization of novel solid electrolytes and their applications in solid state battery systems, fuel cells and sensors. [2]

The term solid state ionics was coined in 1967 by Takehiko Takahashi, [3] but did not become widely used until the 1980s, with the emergence of the journal Solid State Ionics . The first international conference on this topic was held in 1972 in Belgirate, Italy, under the name "Fast Ion Transport in Solids, Solid State Batteries and Devices". [2]

History

Foundations

Michael Faraday in 1842 M Faraday Th Phillips oil 1842.jpg
Michael Faraday in 1842

In the early 1830s, Michael Faraday laid the foundations of electrochemistry and solid-state ionics by discovering the motion of ions in liquid and solid electrolytes. Earlier, around 1800, Alessandro Volta used a liquid electrolyte in his voltaic pile, the first electrochemical battery, but failed to realize that ions are involved in the process. Meanwhile, in his work on decomposition of solutions by electric current, Faraday used not only the ideas of ion, cation, anion, electrode, anode, cathode, electrolyte and electrolysis, but even the present-day terms for them. [4] [5] Faraday associated electric current in an electrolyte with the motion of ions, and discovered that ions can exchange their charges with an electrode while they were transformed into elements by electrolysis. He quantified those processes by two laws of electrolysis. The first law (1832) stated that the mass of a product at the electrode, Δm, increases linearly with the amount of charge passed through the electrolyte, Δq. The second law (1833) established the proportionality between Δm and the “electrochemical equivalent” and defined the Faraday constant F as F = (Δq/Δm)(M/z), where M is the molar mass and z is the charge of the ion.

In 1834, Faraday discovered ionic conductivity in heated solid electrolytes Ag2S and PbF2. [4] In PbF2, the conductivity increase upon heating was not sudden, but spread over a hundred degrees Celsius. Such behavior, called Faraday transition, [6] is observed in the cation conductors Na2S and Li4SiO4 and anion conductors PbF2, CaF2, SrF2, SrCl2 and LaF3. [2]

Later in 1891, Johann Wilhelm Hittorf reported on the ion transport numbers in electrochemical cells, [7] and in the early 20th century those numbers were determined for solid electrolytes. [8]

First theories and applications

The voltaic pile stimulated a series of improved batteries, such as the Daniell cell, fuel cell and lead acid battery. Their operation was largely understood in the late 1800s from the theories by Wilhelm Ostwald and Walther Nernst. In 1894 Ostwald explained the energy conversion in a fuel cell and stressed that its efficiency was not limited by thermodynamics. [9] Ostwald, together with Jacobus Henricus van 't Hoff, and Svante Arrhenius, was a founding father of electrochemistry and chemical ionic theory, and received a Nobel prize in chemistry in 1909.

His work was continued by Walther Nernst, who derived the Nernst equation and described ionic conduction in heterovalently doped zirconia, which he used in his Nernst lamp. Nernst was inspired by the dissociation theory of Arrhenius published in 1887, which relied on ions in solution. [10] In 1889 he realized the similarity between electrochemical and chemical equilibria, and formulated his equation that correctly predicted the output voltage of various electrochemical cells based on liquid electrolytes from the thermodynamic properties of their components. [11]

Besides his theoretical work, in 1897 Nernst patented the first lamp that used a solid electrolyte. [12] Contrary to the existing carbon-filament lamps, Nernst lamp could operate in air and was twice more efficient as its emission spectrum was closer to that of daylight. AEG, a lighting company in Berlin, bought the Nernst’s patent for one million German gold marks, which was a fortune at the time, and used 800 of Nernst lamps to illuminate their booth at the world’s fair Exposition Universelle (1900). [2]

Ionic conductivity in silver halides

Temperature-dependent ionic conductivity of silver halides, original graph by Tubandt and Lorenz. Ionic conductivity of silver halides.png
Temperature-dependent ionic conductivity of silver halides, original graph by Tubandt and Lorenz.

Among several solid electrolytes described in the 19th and early 20th century, α-AgI, the high-temperature crystalline form of silver iodide, is widely regarded as the most important one. Its electrical conduction was characterized by Carl Tubandt and E. Lorenz in 1914. [13] Their comparative study of AgI, AgCl and AgBr demonstrated that α-AgI, is thermally stable and highly conductive between 147 and 555 °C; the conductivity weakly increased with temperature in this range and then dropped upon melting. This behavior was fully reversible and excluded non-equilibrium effects. Tubandt and Lorenz described other materials with a similar behavior, such as α-CuI, α-CuBr, β-CuBr, and high-temperature phases of Ag2S, Ag2Se and Ag2Te. [14] They associated the conductivity with cations in silver and cuprous halides and with ions and electrons in silver chalcogenides.

Point defects in ionic crystals

Frenkel defect in AgCl Frenkel defect in AgCl.png
Frenkel defect in AgCl

In 1926, Yakov Frenkel suggested that in an ionic crystal like AgI, in thermodynamic equilibrium, a small fraction of the cations, α, are displaced from their regular lattice sites into interstitial positions. [15] He related α with the Gibbs energy for the formation of one mol of Frenkel pairs, ΔG, as α = exp(-ΔG/2RT), where T is temperature and R is the gas constant; for a typical value of ΔG = 100 kJ/mol, α ~ 1×10−6 at 100 °C and ~6×10−4 at 400 °C. This idea naturally explained the presence of an appreciable fraction of mobile ions in otherwise defect-free ionic crystals, and thus the ionic conductivity in them. [2]

Frenkel’s idea was expanded by Carl Wagner and Walter Schottky in their 1929 theory, which described the equilibrium thermodynamics of point defects in ionic crystals. In particular, Wagner and Schottky related the deviations from stoichiometry in those crystals with the chemical potentials of the crystal components, and explained the phenomenon of mixed electronic and ionic conduction. [16] [17]

Wagner and Schottky considered four extreme cases of point-defect disorder in a stoichiometric binary ionic crystal of type AB: [17]

  1. Pairs of interstitial cations A+ and lattice vacancies (Frenkel defects)
  2. Pairs of interstitial anions B and lattice vacancies (anti-Frenkel defects)
  3. Pairs of interstitial cations A+ and interstitial anions B with no vacancies
  4. Pairs of A and B-type lattice vacancies with no interstitials (Schottky disorder).

Type-3 disorder does not occur in practice, and type 2 is observed only in rare cases when anions are smaller than cations, while both types 1 and 4 are common and show the same exp(-ΔG/2RT) temperature dependence. [2]

Later in 1933, Wagner suggested that in metal oxides an excess of metal would result in extra electrons, while a deficit of metal would produce electron holes, i.e., that atomic non-stoichiometry would result in a mixed ionic-electronic conduction. [18]

Other types of disorder

Ionic glasses

The studies of crystalline ionic conductors where excess ions were provided by point defect continued through 1950s, and the specific mechanism of conduction was established for each compound depending on its ionic structure. The emergence of glassy and polymeric electrolytes in the late 1970s provided new ionic conduction mechanisms. A relatively wide range of conductivities was attained in glasses, wherein mobile ions were dynamically decoupled from the matrix. [19] It was found that the conductivity could be increased by doping a glass with certain salts, or by using a glass mixture. The conductivity values could be as high as 0.03 S/cm at room temperature, with activation energies as low as 20 kJ/mol. [20] Compared to crystals, glasses have isotropic properties, continuously tunable composition and good workability; they lack the detrimental grain boundaries and can be molded into any shape, but understanding their ionic transport was complicated by the lack of long-range order. [2]

Historically, an evidence for ionic conductivity was provided back in the 1880s, when German scientists noticed that a well-calibrated thermometer made of Thuringian glass would show −0.5 °C instead of 0 °C when placed in ice shortly after immersion in boiling water, and recover only after several months. In 1883, they reduced this effect 10 times by replacing a mixture of sodium and potassium in the glass by either sodium or potassium. [21] This finding helped Otto Schott develop the first accurate lithium-based thermometer. More systematic studies on ionic conductivity in glass appeared in 1884, [22] but received broad attention only a century later. Several universal laws have been empirically formulated for ionic glasses and extended to other ionic conductors, such as the frequency dependence of electrical conductivity σ(ν) – σ(0) ~ νp, where the exponent p depends on the material, but not on temperature, at least below ~100 K. This behavior is a fingerprint of activated hopping conduction among nearby sites. [2]

Polymer electrolytes

In 1975, Peter V. Wright, a polymer chemist from Sheffield (UK), produced the first polymer electrolyte, which contained sodium and potassium salts in a polyethylene oxide (PEO) matrix. [23] Later another type of polymer electrolytes, polyelectrolyte, was put forward, where ions moved through an electrically charged, rather than neutral, polymer matrix. Polymer electrolytes showed lower conductivities than glasses, but they were cheaper, much more flexible and could be easier machined and shaped into various forms. [24] While ionic glasses are typically operated below, polymer conductors are typically heated above their glass transition temperatures. Consequently, both the electric field and mechanical deformation decay on a similar time scale in polymers, but not in glasses. [19] [24] Between 1983 and 2001 it was believed that the amorphous fraction is responsible for ionic conductivity, i.e., that (nearly) complete structural disorder is essential for the fast ionic transport in polymers. [19] However, a number of crystalline polymers have been described in 2001 and later with ionic conductivity as high as 0.01 S/cm 30 °C and activation energy of only 0.24 eV. [2]

Nanostructures

In the 1970s–80s, it was realized that nanosized systems may affect ionic conductivity, opening a new field of nanoionics. In 1973, it was reported that ionic conductivity of lithium iodide (LiI) crystals could be increased 50 times by adding to it a fine powder of ‘’insulating’’ material (alumina). [25] This effect was reproduced in the 1980s in Ag- and Tl-halides doped with alumina nanoparticles. [26] [27] [28] Similarly, addition of insulating nanoparticles helped increase the conductivity of ionic polymers. [29] [30] These unexpected results were explained by charge separation at the matrix-nanoparticle interface that provided additional conductive channels to the matrix, and the small size of the filler particles was required to increase the area of this interface. [26] Similar charge-separation effects were observed for grain boundaries in crystalline ionic conductors. [2]

Applications

By 1971, solid-state cells and batteries based on rubidium silver iodide (RbAg4I5) have been designed and tested in a wide range of temperatures and discharge currents. [31] Despite the relatively high conductivity of RbAg4I5, they have never been commercialized due to a low overall energy content per unit weight (ca. 5 W·h/kg). [32] On the contrary, LiI, which had a conductivity of only ca. 1×10−7 S/cm at room temperature, found a wide-scale application in batteries for artificial pacemakers. The first such device, based on undoped LiI, was implanted into a human in March 1972 in Ferrara, Italy. [33] Later models used as electrolyte a film of LiI, which was doped with alumina nanoparticles to increase its conductivity. [25] LiI was formed in an in situ chemical reaction between the Li anode and iodine-poly(2-vinylpyridine) cathode, and therefore was self-healed from erosion and cracks during the operation. [34]

Sodium-sulfur cells, based on ceramic β-Al2O3 electrolyte sandwiched between molten-sodium anode and molten-sulfur cathode showed high energy densities and were considered for car batteries in the 1990s, but disregarded due to the brittleness of alumina, which resulted in cracks and critical failure due to reaction between molten sodium and sulfur. Replacement of β-Al2O3 with NASICON did not save this application because it did not solve the cracking problem, and because NASICON reacted with the molten sodium. [2]

Yttria-stabilized zirconia is used as a solid electrolyte in oxygen sensors in cars, generating voltage that depends on the ratio of oxygen and exhaust gas and providing electronic feedback to the fuel injector. [35] Such sensors are also installed at many metallurgical and glass-making factories. [36] Similar sensors of CO2, chlorine and other gases based on solid silver halide electrolytes have been proposed in the 1980s–1990s. [2] Since mid-1980s, a Li-based solid electrolyte is used to separate the electrochromic film (typically WO3) and ion-storing film (typically LiCoO2) in the smart glass, [37] a window whose transparency is controlled by external voltage. [38]

Solid-state ionic conductors are essential components of lithium-ion batteries, proton exchange membrane fuel cells (PEMFCs), supercapacitors, a novel class of electrochemical energy storage devices, and solid oxide fuel cells, devices that produces electricity from oxidizing a fuel. Nafion, a flexible fluoropolymer-copolymer discovered in the late 1960s, is widely used as a polymer electrolyte in PEMFCs. [2]

See also

Related Research Articles

An electrolyte is a medium containing ions that are electrically conductive through the movement of those ions, but not conducting electrons. This includes most soluble salts, acids, and bases dissolved in a polar solvent, such as water. Upon dissolving, the substance separates into cations and anions, which disperse uniformly throughout the solvent. Solid-state electrolytes also exist. In medicine and sometimes in chemistry, the term electrolyte refers to the substance that is dissolved.

<span class="mw-page-title-main">Lithium polymer battery</span> Lithium-ion battery using a polymer electrolyte

A lithium polymer battery, or more correctly lithium-ion polymer battery, is a rechargeable battery of lithium-ion technology using a polymer electrolyte instead of a liquid electrolyte. Highly conductive semisolid (gel) polymers form this electrolyte. These batteries provide higher specific energy than other lithium battery types and are used in applications where weight is a critical feature, such as mobile devices, radio-controlled aircraft and some electric vehicles.

<span class="mw-page-title-main">Proton conductor</span> Type of electrolyte

A proton conductor is an electrolyte, typically a solid electrolyte, in which H+ are the primary charge carriers.

Nanoionics is the study and application of phenomena, properties, effects, methods and mechanisms of processes connected with fast ion transport (FIT) in all-solid-state nanoscale systems. The topics of interest include fundamental properties of oxide ceramics at nanometer length scales, and fast ion conductor /electronic conductor heterostructures. Potential applications are in electrochemical devices for conversion and storage of energy, charge and information. The term and conception of nanoionics were first introduced by A.L. Despotuli and V.I. Nikolaichik in January 1992.

Beta-alumina solid electrolyte (BASE) is a fast ion conductor material used as a membrane in several types of molten salt electrochemical cell. Currently there is no known substitute available. β-Alumina exhibits an unusual layered crystal structure which enables very fast ion transport. β-Alumina is not an isomorphic form of aluminium oxide (Al2O3), but a sodium polyaluminate. It is a hard polycrystalline ceramic, which, when prepared as an electrolyte, is complexed with a mobile ion, such as Na+, K+, Li+, Ag+, H+, Pb2+, Sr2+ or Ba2+ depending on the application. β-Alumina is a good conductor of its mobile ion yet allows no non-ionic (i.e., electronic) conductivity. The crystal structure of the β-alumina provides an essential rigid framework with channels along which the ionic species of the solid can migrate. Ion transport involves hopping from site to site along these channels. Since the 1970's this technology has been thoroughly developed, resulting in interesting applications. Its special characteristics on ion and electrical conductivity make this material extremely interesting in the field of energy storage.

<span class="mw-page-title-main">Fast-ion conductor</span>

In materials science, fast ion conductors are solid conductors with highly mobile ions. These materials are important in the area of solid state ionics, and are also known as solid electrolytes and superionic conductors. These materials are useful in batteries and various sensors. Fast ion conductors are used primarily in solid oxide fuel cells. As solid electrolytes they allow the movement of ions without the need for a liquid or soft membrane separating the electrodes. The phenomenon relies on the hopping of ions through an otherwise rigid crystal structure.

The electrochemical window (EW) of a substance is the electrode electric potential range between which the substance is neither oxidized nor reduced. The EW is one of the most important characteristics to be identified for solvents and electrolytes used in electrochemical applications. The EW is a term that is commonly used to indicate the potential range and the potential difference. It is calculated by subtracting the reduction potential from the oxidation potential.

<span class="mw-page-title-main">Ionic conductivity (solid state)</span>

Ionic conductivity is a measure of a substance's tendency towards ionic conduction. Ionic conduction is the movement of ions. The phenomenon is observed in solids and solutions. Ionic conduction is one mechanism of current.

An advanced superionic conductor (AdSIC) in materials science, is fast ion conductor that has a crystal structure close to optimal for fast ion transport (FIT).

<span class="mw-page-title-main">Yttria-stabilized zirconia</span> Ceramic with room temperature stable cubic crystal structure

Yttria-stabilized zirconia (YSZ) is a ceramic in which the cubic crystal structure of zirconium dioxide is made stable at room temperature by an addition of yttrium oxide. These oxides are commonly called "zirconia" (ZrO2) and "yttria" (Y2O3), hence the name.

<span class="mw-page-title-main">Solid-state battery</span> Battery with solid electrodes and a solid electrolyte

A solid-state battery is an electrical battery that uses a solid electrolyte for ionic conductions between the electrodes, instead of the liquid or gel polymer electrolytes found in conventional batteries. Solid-state batteries theoretically offer much higher energy density than the typical lithium-ion or lithium polymer batteries.

The lithium–air battery (Li–air) is a metal–air electrochemical cell or battery chemistry that uses oxidation of lithium at the anode and reduction of oxygen at the cathode to induce a current flow.

LISICON is an acronym for LIthiumSuper Ionic CONductor, which refers to a family of solids with the chemical formula Li2+2xZn1−xGeO4.

<span class="mw-page-title-main">NASICON</span> Class of solid materials

NASICON is an acronym for sodium (Na) super ionic conductor, which usually refers to a family of solids with the chemical formula Na1+xZr2SixP3−xO12, 0 < x < 3. In a broader sense, it is also used for similar compounds where Na, Zr and/or Si are replaced by isovalent elements. NASICON compounds have high ionic conductivities, on the order of 10−3 S/cm, which rival those of liquid electrolytes. They are caused by hopping of Na ions among interstitial sites of the NASICON crystal lattice.

<span class="mw-page-title-main">Mixed conductor</span>

Mixed conductors, also known as mixed ion-electron conductors(MIEC), are a single-phase material that has significant conduction ionically and electronically. Due to the mixed conduction, a formally neutral species can transport in a solid and therefore mass storage and redistribution are enabled. Mixed conductors are well known in conjugation with high-temperature superconductivity and are able to capacitate rapid solid-state reactions.

<span class="mw-page-title-main">Solid-state electrolyte</span> Type of solid ionic conductor electrolyte

A solid-state electrolyte (SSE) is a solid ionic conductor and electron-insulating material and it is the characteristic component of the solid-state battery. It is useful for applications in electrical energy storage (EES) in substitution of the liquid electrolytes found in particular in lithium-ion battery. The main advantages are the absolute safety, no issues of leakages of toxic organic solvents, low flammability, non-volatility, mechanical and thermal stability, easy processability, low self-discharge, higher achievable power density and cyclability. This makes possible, for example, the use of a lithium metal anode in a practical device, without the intrinsic limitations of a liquid electrolyte thanks to the property of lithium dendrite suppression in the presence of a solid-state electrolyte membrane. The use of a high capacity anode and low reduction potential, like lithium with a specific capacity of 3860 mAh g−1 and a reduction potential of -3.04 V vs SHE, in substitution of the traditional low capacity graphite, which exhibits a theoretical capacity of 372 mAh g−1 in its fully lithiated state of LiC6, is the first step in the realization of a lighter, thinner and cheaper rechargeable battery. Moreover, this allows the reach of gravimetric and volumetric energy densities, high enough to achieve 500 miles per single charge in an electric vehicle. Despite the promising advantages, there are still many limitations that are hindering the transition of SSEs from academia research to large-scale production, depending mainly on the poor ionic conductivity compared to that of liquid counterparts. However, many car OEMs (Toyota, BMW, Honda, Hyundai) expect to integrate these systems into viable devices and to commercialize solid-state battery-based electric vehicles by 2025.

Lithium lanthanum zirconium oxide (LLZO, Li7La3Zr2O12) or lithium lanthanum zirconate is a lithium-stuffed garnet material that is under investigation for its use in solid-state electrolytes in lithium-based battery technologies. LLZO has a high ionic conductivity and thermal and chemical stability against reactions with prospective electrode materials, mainly lithium metal, giving it an advantage for use as an electrolyte in solid-state batteries. LLZO exhibits favorable characteristics, including the accessibility of starting materials, cost-effectiveness, and straightforward preparation and densification processes. These attributes position this zirconium-containing lithium garnet as a promising solid electrolyte for all-solid-state lithium-ion rechargeable batteries.

A polymer electrolyte is a polymer matrix capable of ion conduction. Much like other types of electrolyte—liquid and solid-state—polymer electrolytes aid in movement of charge between the anode and cathode of a cell. The use of polymers as an electrolyte was first demonstrated using dye-sensitized solar cells. The field has expanded since and is now primarily focused on the development of polymer electrolytes with applications in batteries, fuel cells, and membranes.

<span class="mw-page-title-main">Lithium aluminium germanium phosphate</span> Chemical compound

Lithium aluminium germanium phosphate, typically known with the acronyms LAGP or LAGPO, is an inorganic ceramic solid material whose general formula is Li
1+x
Al
x
Ge
2-x
(PO
4
)
3
. LAGP belongs to the NASICON family of solid conductors and has been applied as a solid electrolyte in all-solid-state lithium-ion batteries. Typical values of ionic conductivity in LAGP at room temperature are in the range of 10–5 - 10–4 S/cm, even if the actual value of conductivity is strongly affected by stoichiometry, microstructure, and synthesis conditions. Compared to lithium aluminium titanium phosphate (LATP), which is another phosphate-based lithium solid conductor, the absence of titanium in LAGP improves its stability towards lithium metal. In addition, phosphate-based solid electrolytes have superior stability against moisture and oxygen compared to sulfide-based electrolytes like Li
10
GeP
2
S
12
(LGPS) and can be handled safely in air, thus simplifying the manufacture process. Since the best performances are encountered when the stoichiometric value of x is 0.5, the acronym LAGP usually indicates the particular composition of Li
1.5
Al
0.5
Ge
1.5
(PO
4
)
3
, which is also the typically used material in battery applications.

Fluoride batteries are rechargeable battery technology based on the shuttle of fluoride, the anion of fluorine, as ionic charge carriers.

References

  1. Chowdari, B. V. R. (2004). Proceedings of the 9th Asian Conference on Solid State Ionics the science and technology of ions in motion: Jeju Island, South Korea, 6–11 June 2004. Singapore River Edge, NJ: World Scientific. ISBN   9789812702586.
  2. 1 2 3 4 5 6 7 8 9 10 11 12 13 Funke, K. (2013). "Solid State Ionics: From Michael Faraday to green energy—the European dimension". Science and Technology of Advanced Materials. 14 (4): 043502. Bibcode:2013STAdM..14d3502F. doi:10.1088/1468-6996/14/4/043502. PMC   5090311 . PMID   27877585.
  3. Yamamoto, Osamu (2017). "Solid state ionics: A Japan perspective". Science and Technology of Advanced Materials. 18 (1): 504–527. Bibcode:2017STAdM..18..504Y. doi:10.1080/14686996.2017.1328955. PMC   5532972 . PMID   28804526.
  4. 1 2 Faraday, M. (1839) Experimental Researches in Electricity, Art. 1339, Taylor and Francis, London.
  5. See Harper, Douglas. "ion". Online Etymology Dictionary . and other OED pages for the etymology of these terms
  6. O’Keeffe, M. (1976). Mahan, G. D.; Roth, W. L. (eds.). Superionic Conductors. New York: Plenum Press. p. 101. doi:10.1007/978-1-4615-8789-7_9. ISBN   978-1-4615-8791-0.
  7. Hittorf, J.W. (1892). Z. Phys. Chem. 10: 593.{{cite journal}}: Missing or empty |title= (help)
  8. Tubandt, C. (1921). "Über Elektrizitätsleitung in festen kristallisierten Verbindungen. Zweite Mitteilung. Überführung und Wanderung der Ionen in einheitlichen festen Elektrolyten". Zeitschrift für anorganische und allgemeine Chemie. 115: 105–126. doi:10.1002/zaac.19211150106.
  9. Ostwald (1894). "Zeitschrift für Elektrotechnik und Elektrochemie. Die Wissenschaftliche Elektrochemie der Gegenwart und die Technische der Zukunft". Zeitschrift für Elektrotechnik und Elektrochemie. 1 (4): 122–125. doi:10.1002/bbpc.18940010403.
  10. Arrhenius, S. (1887). "Über die Dissociation der in Wasser gelösten Stoffe". Z. Phys. Chem. 1: 631–648. doi:10.1515/zpch-1887-0164. S2CID   102373219.
  11. Nernst, W. (1926) Theoretische Chemie, Enke, Stuttgart
  12. Nernst, W. (1899) pp. 192 and 367 in Mutter Erde, Spemann, Berlin, vol. 2.
  13. 1 2 Tubandt, C.; Lorenz, E. (1914). "Molekularzustand und elektrisches Leitvermögen kristallisierter Salze". Z. Phys. Chem. B. 24: 513–543. doi:10.1515/zpch-1914-8737. S2CID   99214772.
  14. Tubandt, C. (1932) in: Handbuch der Experimentalphysik XII, part 1, W. Wien and F. Harms (eds.), Akadem. Verlagsges., Leipzig.
  15. Frenkel, J. (1926). "Über die Wärmebewegung in festen und flüssigen Körpern". Zeitschrift für Physik. 35 (8–9): 652–669. Bibcode:1926ZPhy...35..652F. doi:10.1007/BF01379812. S2CID   121391169.
  16. Schottky, W.; Ulich, H. and Wagner, C. (1929) Thermodynamik, Springer, Berlin.
  17. 1 2 Wagner, C.; Schottky, W. (1930). "Theorie der geordneten Mischphasen" [Theory of arranged mixed phases]. Z. Phys. Chem. B. 11: 163.
  18. Wagner C (1933). "Theorie der geordneten Mischphasen. III. Felordnungserscheinungen in polaren Verbindungen als Grundlage für Ionen- und Elektronenleitung" [Theory of arranged mixed phases. III. Disarranged phenomena in polar compounds as basis for ionic and electronic conduction]. Z. Phys. Chem. B. 22: 181–194. doi:10.1515/zpch-1933-2213. S2CID   202044725.
  19. 1 2 3 Angell, C. (1983). "Fast ion motion in glassy and amorphous materials". Solid State Ionics. 9–10: 3–16. doi:10.1016/0167-2738(83)90206-0.
  20. Magistris, A.; Chiodelli, G.; Schiraldi, A. (1979). "Formation of high conductivity glasses in the system AgI-Ag2O-B2O3". Electrochimica Acta. 24 (2): 203. doi:10.1016/0013-4686(79)80025-0.
  21. Weber R. (1883) Berliner Akad. Wiss. II 1233
  22. Warburg, E. (1884). "Ueber die Electrolyse des festen Glases". Annalen der Physik. 257 (4): 622–646. Bibcode:1884AnP...257..622W. doi:10.1002/andp.18832570406.
  23. Wright, P. V. (1975). "Electrical conductivity in ionic complexes of poly(ethylene oxide)". British Polymer Journal. 7 (5): 319–327. doi:10.1002/pi.4980070505.
  24. 1 2 Armand, M. (1983). "Polymer solid electrolytes – an overview". Solid State Ionics. 9–10: 745–754. doi:10.1016/0167-2738(83)90083-8.
  25. 1 2 Liang, C. C. (1973). "Conduction Characteristics of the Lithium Iodide-Aluminum Oxide Solid Electrolytes". Journal of the Electrochemical Society. 120 (10): 1289. Bibcode:1973JElS..120.1289L. doi: 10.1149/1.2403248 .
  26. 1 2 Maier, J. (1987). "Defect Chemistry and Conductivity Effects in Heterogeneous Solid Electrolytes". Journal of the Electrochemical Society. 134 (6): 1524–1535. Bibcode:1987JElS..134.1524M. doi:10.1149/1.2100703.
  27. Maier, J.; Reichert, B. (1986). "Ionic Transport in Heterogeneously and Homogeneously Doped Thallium (I)-Chloride". Berichte der Bunsengesellschaft für physikalische Chemie. 90 (8): 666. doi:10.1002/bbpc.19860900809.
  28. Shahi, K.; Wagner, J. B. (1980). "Fast ion transport in silver halide solid solutions and multiphase systems". Applied Physics Letters. 37 (8): 757. Bibcode:1980ApPhL..37..757S. doi:10.1063/1.92023.
  29. Wieczorek, W.; Such, K.; Przyłuski, J.; Floriańczyk, Z. (1991). "Blend-based and composite polymer solid electrolytes". Synthetic Metals. 45 (3): 373. doi:10.1016/0379-6779(91)91792-9.
  30. Scrosati, B.; Croce, F.; Appetecchi, G. B.; Persi, L. (1998). "Nanocomposite polymer electrolytes for lithium batteries". Nature. 394 (6692): 456. Bibcode:1998Natur.394..456C. doi:10.1038/28818. S2CID   4368681.
  31. Owens B. B. (1971) Advances in Electrochemistry and Electrochemical Engineering. vol 8. P. Delahay and C. W. Tobias (eds.). New York: Wiley-Interscience. p. 1. ISBN   0471875260.
  32. Yamamoto, O. (1995). Bruce, P. G. (ed.). Solid State Electrochemistry. Cambridge: Cambridge University Press. p. 292. ISBN   0521599490.
  33. Owens, B. B. (2000). "Solid state electrolytes: Overview of materials and applications during the last third of the Twentieth Century". Journal of Power Sources. 90 (1): 2–8. Bibcode:2000JPS....90....2O. doi:10.1016/S0378-7753(00)00436-5.
  34. Owens, B. B.; Oxley, J. E.; Sammels, A. F. (1977). Geller, S. (ed.). Solid Electrolytes. Berlin: Springer. p.  67. doi:10.1007/3540083383_4. ISBN   978-3-540-08338-2.
  35. Knauth, P.; Tuller, H. L. (2004). "Solid-State Ionics: Roots, Status, and Future Prospects". Journal of the American Ceramic Society. 85 (7): 1654. doi:10.1111/j.1151-2916.2002.tb00334.x.
  36. Fischer W. A. and Janke D. (1975) Metallurgische Elektrochemie. Berlin: Springer.
  37. Svensson, J. S. E. M.; Granqvist, C. G. (1985). "Electrochromic coatings for "smart windows"". Solar Energy Materials. 12 (6): 391. doi:10.1016/0165-1633(85)90033-4.
  38. Granqvist, C. G. (2008). "Smart Windows". Advances in Science and Technology. Smart Optics. 55: 205–212. doi:10.4028/www.scientific.net/AST.55.205. ISBN   978-3-03813-226-4. S2CID   212748428.