Next Article in Journal
Algicidal Bacteria: A Sustainable Proposal to Control Microalgae in the Conservation and Restoration of Stone Cultural Heritage
Previous Article in Journal
Tannin Supplementation Alters Foraging Behavior and Spatial Distribution in Beef Cattle
Previous Article in Special Issue
Bayesian–Kalman Fusion Framework for Thermal Fault Diagnosis of Battery Energy Storage Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tetracyanoquinodimethane and Its Derivatives as Promising Sustainable Materials for Clean Energy Storage and Conversion Technologies: A Review

by
Tetiana Starodub
Institute of Chemistry, Jan Kochanowski University, Uniwersytecka St. 7G, PL-25406 Kielce, Poland
Sustainability 2025, 17(23), 10612; https://doi.org/10.3390/su172310612
Submission received: 10 September 2025 / Revised: 5 November 2025 / Accepted: 24 November 2025 / Published: 26 November 2025
(This article belongs to the Special Issue Advances in Energy Storage Technologies to Meet Future Energy Demands)

Abstract

7,7′,8,8′-tetracyanoquinodimethane (TCNQ) is one of the most widely studied redox-active molecules and effective π-acceptors, possessing excellent electrical properties. TCNQ derivatives are considered exciting materials due to their multitude of uses. TCNQ-based electrode and dopant materials offer enormous potential in revolutionary energy storage and conversion devices because of their safety, long-term stability, environmental friendliness, charge mobility, and low electrode potential. This review presents the most important advances in applications of TCNQ and its derivatives as promising sustainable organic materials for the optimization of electrical properties in energy storage and conversion devices compared to inorganic materials. This paper shows that materials based on TCNQ can be used as promising candidates for designing efficient and stable photovoltaics, sustainable batteries, and triboelectric nanogenerators, which are very important to the development of energy harvesting. The results presented in this review are sufficiently convincing for us to conclude that applying the TCNQ doping strategy to improve the performance of the discussed devices is very promising. The presented data could serve as inspiration for the development and subsequent design of new, effective, high-performance micro- and nano-systems based on TCNQ for energy storage and conversion.

1. Introduction

Society’s increasing demand for energy is one of the greatest challenges for humanity. Therefore, the most important goal for today’s world is to find renewable, sustainable and ecological sources of energy. Current research in the field of energy covers topics such as the acquisition of energy, sustainable energy sources for micro- and nano-systems, the storage of energy, energy conversion, renewable energy, and green energy. An important aspect of the development of energy harvesting is also achieving it on a smaller scale for electrodes, sensitive chemical and biological sensors, electromechanical systems, nanorobots, and portable personal electronics. The development of processing technologies and advanced electrode materials to satisfy requirements relating to reliability and cost-effectiveness is important as well.
Since the 1960s, the production and application of 7,7′,8,8′-tetracyanoquinodimethane and its derivatives have been actively developing [1,2,3,4]. TCNQ has been one of the most commonly studied redox-active molecules since its first synthesis in 1962 [5]. The body of research on the use of this molecule is constantly growing [1,2,3,4,5,6,7,8,9]. TCNQ has a central hexagonal ring, which is aromatized when taking an electron, and four low-lying vacant C≡N π*-orbitals. This can help to accommodate and delocalize the incoming electron [10]. Moreover, TCNQ, as an active acceptor molecule and redox ligand, can easily transform into an open-shell TCNQ•− radical anion via electrochemical methods or chemical reduction when it is placed in contact with different electron donors (Figure 1) [2,4,11,12]. The morphology of TCNQ depends on its form and its interaction with the surface of other materials. As a result, different structures are formed [2,5,6]. This variability is a result of different synthesis methods and conditions. The neutral form of TCNQ is often grown as a thin film on substrates, while TCNQ salts and complexes can be grown as microcrystals or nanostructures. However, it should be taken into account that the morphology of TCNQ structures is dependent on specific reactions, such as vapor transport, and the composition of the system. This is related to additional requirements when constructing devices.
The four widely separated cyan groups (–CN) in the TCNQ structure may also act as coordination sites, making TCNQ a potential polydentate ligand capable of binding to four metal centers [1,6,13]. After reacting with a metal ion, including the charge transfer from a metal ion to the TCNQ molecule, TCNQ forms a charge transfer complex (CTC) which participates in the formation of a close π–π bond. This π stacked column structure becomes the basis for the enviable electric and magnetic properties of TCNQ-based compounds [2,4,14]. Charge transfer in TCNQ involves the transfer of an electron from a donor molecule to an electron-accepting TCNQ molecule, forming highly conductive salts and CTCs. This interaction between the donor and acceptor is crucial for creating materials with unique electronic properties. Depending on the donor and reaction conditions, TCNQ-based materials can exhibit incomplete or complete charge transfer. The tilt angle between the donor and strong TCNQ acceptor also changes the degree of charge transfer (q) and electrical conductivity. If 0.5 < q < 0.74, the complex usually exhibits metallic behavior because of its segregated structure. Therefore, combinations of different donors with a strong TCNQ acceptor are already successfully used in the electronics industry [2,3,4,14,15,16,17,18,19,20,21,22,23,24,25,26], including spintronic and non-linear optics [2,3,15], biosensing [4,20,21,22], and electrochemical analysis [4,23,24,25,26]. So far, a large number of TCNQ derivatives have been synthesized, exhibiting various unique physicochemical properties [2,3,4,27,28,29,30]. The deposited active layers doped with a TCNQ thin film work as charge transfer electron-accepting layers. According to dimensionality, these layers can be categorized as 0D to 3D [2,3,30].
Converting dispersed mechanical energy into electrical energy can effectively mitigate the global energy shortage problem. Not all current battery technologies meet the needs of the modern world [31,32,33]. The popularity of organic electrode materials based on organic compounds with redox activity as an alternative to conventional inorganic electrodes is increasing due to their safety, stability, and environmental friendliness [34,35,36,37,38]. In particular, since the beginning of the 21st century, TCNQ and its derivatives have demonstrated extraordinary potential for use in different electronics because of their suitability in energy devices [2,4,39,40,41,42]. For example, the small TCNQ molecule can be used as a cathode material for aqueous zinc batteries due to its very low water solubility and energy gap of 2.58 eV [2,4,43,44,45]. This significantly smaller energy gap than most organic electrodes indicates better internal electrical conductivity [46,47]. Therefore, TCNQ could be a promising alternative to standard conductive polymeric materials such as polythiophene (PTh), polypyrrole, and polyaniline (PANI) used in electronic technologies [48,49]. TCNQ-based electrode materials have also been applied in non-aqueous, redox flow, multivalent-metal, dual-ion, and all-solid-state batteries due to their low cost and high sustainability, among other factors [50,51,52,53,54,55]. Thus, TCNQ can clearly be seen as a future alternative for sustainable battery technologies [56]. Scheme 1 summarizes the physicochemical properties of the TCNQ molecule, indicating that TCNQ is a potential candidate for battery construction.
Thin-film solar cells are another important source of renewable energy that is worth mentioning. The most efficient organic solar cells (OSCs) and perovskite solar cells (PSCs) are made from organic materials, including TCNQ thin-film dopants [57,58,59]. These thin films have the potential to reduce the cost of photovoltaics by eliminating the use of silicon wafers, which account for over 50% of the total production cost of solar cell devices. TCNQ doping significantly facilitates the recombination of majority carriers at the OSC interface. Recombination centers are hidden in the case of excitons and minority carriers. This reduces energy losses during electron transfer processes, which leads to higher device efficiency [60]. Because of their solubility in organic solvents, like acetonitrile or chlorobenzene [60,61], TCNQ dopants can also be integrated into fully solution-processed devices. On the other hand, TCNQ-based PSCs stand out for the robustness of their absorber material. TCNQ’s high chemical stability and the wide variety of successful TCNQ preparation methods available make TCNQ suitable for large-scale PSC production. Thus, TCNQ can be used as a “morphology modulator” in organic solar cells. By introducing and then vaporizing TCNQ during thermal annealing, a better-optimized film morphology can be achieved. Recently, organic perovskite photovoltaic cells have shown great promise in energy storage technologies due to their simple and cheap production process and an unprecedentedly fast increase in efficiency [62,63]. TCNQ is a suitable candidate for PSC applications because of its various desirable properties, including its ability to melt without decomposition, variable solid-state morphology, and high conductivity.
In addition to the development of many “green” energy sectors such as wind, water, and solar energy, and the research on methods for storing the energy obtained in these ways, work is also being carried out in parallel on recovering energy generated during everyday activities such as walking or writing. Known for their wide availability and exceptional biocompatibility, TCNQ and some of its derivatives are good candidates for the development of environmentally friendly triboelectric nanogenerators (TENGs) [64]. TCNQ doping helps to reduce mechanical input and surface wear. Energy from such nanogenerators can help to power nano- and micro-devices.
To summarize, thanks to their unique properties, organic TCNQ-based molecular dopants of high-performing and stable solar cells and batteries offer an interesting alternative to inorganic structures. The widespread implementation of new TCNQ dopants not only increases photovoltaic performance but also decreases the optical band gap, enables diffusion-controlled redox processes, and improves manufacturing compatibility (Figure 2).
Furthermore, metal–organic TCNQ-based compounds have also been placed in the spotlight as catalysts, phosphorescent materials, hydrogen storage materials, etc. [65,66]. Some of these materials exhibit metal-like conductivity and superconductivity at low temperatures. Thus, TCNQ and some of its derivatives can be considered alternative sustainable dopants. In this introductory work, for the first time, attention is focused on the recent progress made in three energy technologies based on the use of TCNQ:
  • Modern sustainable TCNQ organic molecular solids for efficient stable photovoltaics;
  • Cathode TCNQ-based materials in sustainable high-capacity and long-lifetime batteries;
  • TCNQ-based triboelectric nanogenerators as an efficient and high-performance wearable electronic power source.
This pioneering review has been prepared with the hope that the data contained herein will help scientists to exploit the full potential of TCNQ-based materials and encourage researchers to continue research in these application areas.

2. Using TCNQ and Its Derivatives for High-Efficiency Stable Photovoltaics

Solar cells are photovoltaic devices that directly convert sunlight into electrical energy [33,67,68,69,70,71]. The history of photovoltaics began with experiments conducted by the French scientist Edmund Becquerel. In 1839, at the age of 19, he discovered the photovoltaic phenomenon—the generation of an electric current under the influence of light. He created the first prototype of a photovoltaic cell, built from two electrodes composed of platinum and silver, immersed in nitric acid. In 1887, German physicist Heinrich Hertz discovered the photoelectric effect, which showed that light could liberate electrons from metal surfaces. This phenomenon became the basis for the further development of photovoltaics, now essential to modern society. The history of photovoltaics, from the scientific discoveries in the 19th century to contemporary innovations, including inventions in the field of TCNQ, is illustrated in Figure 3 [72,73,74,75]. In recent years, rapid developments in the field of photovoltaics have been achieved, both in terms of efficiency and the diversity of technologies [76,77]. Products such as solar roof tiles and transparent panels have started to appear on the market, enabling the increasingly broader use of photovoltaics in urban architecture. Meanwhile, third-generation photovoltaics (PVs), including organic solar cells [23,39,78,79,80], dye-sensitized solar cells (DSSCs) [81,82,83,84,85], solar thermal selective coatings (STSCs) [86], and perovskite solar cells [58,61,62,63,87,88] are also making inroads, opening the door to novel, unconventional PV applications. DSSCs give great hope due to their characteristics such as inexpensive synthesis, simple construction, elasticity, tunable transparency, high efficiency, and eco-friendliness [85,89,90,91]. Meanwhile, PSCs exhibit high potential due to their good optical absorption coefficient, band gap, photovoltaic parameters (open circuit voltage greater than 1.0 V), carrier length, mobility, and conversion efficiency of more than 20%, as compared to silicon and gallium arsenide solar cells [92,93,94,95].
In recent times, organic batteries have become very popular [96,97]. In the research on OSCs [72,75,98,99], modern organic TCNQ-based molecular solids also offer an interesting alternative to inorganic structures. These sustainable TCNQ systems are excellent photovoltaic materials for the production of energy devices, with the ability to obtain a fill factor (FF) and high open-circuit voltage (Voc). In 2012, organic solar cells based on TCNQ/copper phthalocyanine (CuPc) and TCNQ/zinc phthalocyanine (ZnPc) were fabricated [100]. The relationship between their microstructure and photovoltaic properties has been investigated. In the case of a TCNQ/ZnPc heterojunction solar cell, the open-circuit voltage, short circuit current (Jsc), fill factor, and power conversion efficiency (PCE) were found to be 0.58 V, 0.27 × 10−6 A cm−2, 0.18, and 1.6 × 10−5%, respectively. Regarding a TCNQ/CuPc solar cell, the Voc, Jsc, FF, and PCE were obtained to be 0.48 V, 0.37 × 10−6 A cm−2, 0.16, and 2.8 × 10−5%, respectively. According to Suzuki et al., the low performance in both cases is caused by the low crystallinity of the TCNQ layers, which are responsible for the charge transfer from the ZnPc electron-donor layer to the TCNQ acceptor layer [100]. The authors suggest that in both cases of the heterojunction TCNQ/CuPc and TCNQ/ZnPc solar cells, the TCNQ thin film acts as a strong electron-accepting layer and an n-type semiconductor. Ref. [100] is one of the few highly valuable works which propose guidelines for the optimization of photovoltaic properties in heterojunction solar cells.
Large-area graphene is a promising candidate for applications in elastic optoelectronic devices. In 2012, a TCNQ doping process on graphene for the purposes of forming sandwiched graphene/TCNQ/graphene stacked films for solar cell anodes was reported [101]. In this case, TCNQ p-dopants were securely embedded between two graphene layers. The anode based on sandwiched graphene/TCNQ/graphene stacked films showed optimum PCE (~2.58%). This is an encouraging result, because power conversion efficiency using a pristine chemical vapor deposition (CVD) graphene anode is still not appealing due to its much lower conductivity [102]. The TCNQ-based anode film proposed in [101] is propitious for next-generation extensible conductivity devices.
Organic PVs have attracted attention in research [77,97,98,103] due to their high bendability and compatibility with electronic fabrication processes. For example, a very strong electron acceptor tetrafluoro-tetracyanoquinodimethane (F4TCNQ) has been proven to be a perfect candidate for manufacturing effective OSCs. Liu et al. [103] incorporated F4TCNQ into a poly(3-hexylthiophene) (P3HT), leading to the realization of devices with an improved power conversion efficiency of 5.83%. Figure 4 shows an electron transferring from P3HT to F4TCNQ. This TCNQ derivative, with an exceptionally high potential value of 5.2 eV, possesses an even stronger electron-accepting character than TCNQ [104]. Therefore, the excited electrons in P3HT have a high possibility of transferring to F4TCNQ, thus reducing the recombination of electrons and holes. F4TCNQn− (n = 0, 1 or 2) also forms strong intermolecular hydrogen bonds, favoring strong electronic interactions [105,106]. In addition, a fluoride can increase the redox potential of the organic molecule and improve the average operating voltage. As a result, the higher charge carrier mobility and photoconductivity of F4TCNQ/P3HT OSCs with an increased Jsc from 8.19 to 9.85 mA cm−2 and an increased FF from 63.2% to 68.0% is observed. The electron mobility of the F4TCNQ-doped device increased from 4.92 × 10−4 to 6.91 × 10−4 cm2 V−1 s−1, being higher than that of the control device. The incorporation of F4TCNQ can effectively increase the charge carrier density, which is beneficial for improving the FF of solar cells. Thus, doping increases the quantity of charge carriers, leading to a higher conductivity, as also described in [107,108,109,110]. Moreover, organic p-doping is promising for creating sustainable solar cells due to the high absorption and electrical properties of these cells, which in many respects are comparable to those of crystalline semiconductors. For this reason, p-type doping in organic layers with TCNQ is a good way of improving the performance of OSCs.
The TCNQ molecule can exist in a neutral, anion-radical, or dianion form. The negative charge of the molecule arises from the strengthening of bonds between the cyan groups and the surface atoms by means of the transfer of electrons from the surface into the molecule. As a consequence, the degree of electron density (degree of charge transfer) in the TCNQ molecule increases. The bond lengths of –C=C– and –C≡N are reduced, resulting in the formation of a stable TCNQ•− radical anion with an active center in its structure. These active sites help to maintain strong contacts (bonds) with other active molecules. Thanks to this, 7,7′,8,8′-tetracyanoquinodimethane also experiences a nucleophilic substitution reaction when it reacts with primary or secondary amines (Scheme 2) to form a number of diaminodicyanoquinodimethane (DADQs) compounds [60,111,112,113,114,115,116,117] with attractive and unpredictable properties. The ease of formation of these compounds results from TCNQ being one of the strongest acceptors and the amino group (–NH2) being a strong electron donor group. Consequently, TCNQ-based DADQ compounds have been used as conductive materials with promising optoelectronic effects.
For example, in 2023, Mohitkar and his team [60] described an amine-TCNQ/TiO2 solar cell that displays an improved Voc of 3 V, a Jsc from 1.25 to 9.12 mA cm−2, an FF from 0.21 to 0.59, and an increased PCE from 2.26% to 11.75%. The optical band gap among DADQ molecules is about 2.0 eV, which also suggests enhanced intramolecular charge transfer. The authors also concluded that these DADQ compounds have potential applications in OSCs.
TCNQ doping also enhances the performance of dye-sensitized solar cells [118]. Dye sensitizer is the most important part of DSSCs affecting photovoltaic processes [119,120,121]. The TCNQ-based dye structure with TiO2 increases the driving force of electrons without undesirable charge recombination between the oxidized dye and the semiconductor in comparison with standard DSSCs [118,122]. The affinity of the dye for electron transfer to the semiconductor increases due to its lower electrophilicity, which then improves the final efficiency. For example, in [118], Pakravesh et al. described (2-(4-(3,3-dicyano-1-(4-(dimethylamino)-phenyl)allylidene)cyclohexa-2,5-dien-1-ylidene)malononitrile)/TCNQ/TiO2 (AQ) and (2-(4-(3,3-dicyano-1-(4-(diphenylamino)-phenyl)allylidene)-cyclohexa-2,5-dien-1-ylidene)malononitrile)/TCNQ/TiO2 (TQ) dye solar cells. These TCNQ-adduct dyes in DSSCs have a higher electron transfer rate constant, longer excited-state lifetime, and higher polarization. In the case of AQ dye solar cells, the measured parameters Voc, Jsc, and FF were found to be 3.25 V, 10.63 mA cm−2, and 95%, respectively. In the case of TQ dye solar cells, the Voc, Jsc, and FF were 2.72 V, 11.86 mA cm−2, and 94%, respectively. The results obtained allowed the authors to conclude that these AQ and TQ TCNQ-based dyes could be seen as favorable candidates for use in sustainable DSSCs.
In recent years [58,88,92,93,123,124,125,126], there has been an increasing demand for portable energy systems based on embedded perovskite with an excellent power-to-weight ratio, superior efficiency, facile features, and low cost. In single-junction PSCs, a PCE over 25.8% has been realized [127]. Currently, the widespread use of silicon solar cells is limited due to the expensive production process of these devices [128,129]. The manufacturing of sustainable organic–inorganic halide perovskite (OIHP) solar cells [130,131,132,133,134,135] with a high PCE (reaching 26.7%) has emerged as an alternative technology and surprised the community. TCNQ passivation plays a key role in improving the photoelectric properties of perovskite devices. For example, Abicho and co-authors [57] proposed a strategy to introduce a TCNQ interface passivator between OIHP and various organic molecular sheets. The authors showed that TCNQ is an excellent sustainable candidate for minimizing interface defects in OIHP solar cells. Earlier, in 2014 [136], it was also found that large current leakages caused losses in the Voc and FF, which had depended on the quality of the interfaces of the OIHP layers and charge transport. The TCNQ-treated glass/ITO/PEDOT:PSC/OIHP thin films (ITO—indium tin oxide; PEDOT—poly(3,4-ethylenedioxythiophene)) reported by Wang et al. exhibited lower current leakages, which led to increases in Voc by ~1.03 V, Jsc by ~20.1 mA cm− 2, and PCE by 16.5% [136]. The authors therefore concluded that the use of TCNQ interface passivation is a good option to improve the yield of perovskite devices.
One of the most important approaches to improving the efficiency and stability of perovskite solar cells is rational hole transport layer (HTL) material design. Liu and co-workers [137] described the effect of 0.30 wt. % F4TCNQ doping on the optical and electrical properties of these devices. The highest power conversion efficiency was notably improved from 13.30% (undoped) to 17.22% (doped) due to efficient carrier transport from the perovskite absorber sheet to the HTL. The obtained data indicated that the new F4TCNQ-doped film is an efficient HTL, providing high-performance perovskite solar cells. For example, according to [138], the measured FF and PCE parameters for F4TCNQ are 69.9% and 14.3%, respectively. Meanwhile, according to [139], a F4TCNQ interface layer was used to increase efficiency by passivating the halogen bond and doping the interface. In this solar cell, the parameters FF and PCE were measured to be 75.4%, and 16.4, respectively. In contrast, F6TCNQ [140] successfully doped a 2,2′,7,7′-tetrakis[N,N-di(4-methoxyphenyl)amino]-9,9′-spirobifluorene (spiro-OMeTAD) HTL device with a high FF of 79.8% and a higher PCE of 18%.
Zhao and co-workers [141] described a F4TCNQ doping strategy of nickel oxide (NiOx) as a highly efficient hole transporting material for inverted PSCs. Nickel oxide [142,143] is a promising material for hole transport in inverted perovskite solar cells due to its low cost and high photostability. A new high-temperature-resistant F4TCNQ/NiOx film demonstrated reduced surface roughness and a higher carrier transport capacity compared to pure NiOx. The F4TCNQ/NiOx-based hole transport layer exhibits an open-circuit voltage of 1.02 V, a short circuit current density of 20.07 mA cm−2, a fill factor of 74.5%, and a PCE of 15.7%, which is 16.2% higher than that of the undoped HTL. In addition, F4TCNQ-doped NiOx substrates facilitated hole migration to the HTL and therefore were characterized by increased transmittance, improving the final photoelectric properties. These results prove that F4TCNQ/NiOx can be used as an excellent sustainable and controllable hole extraction layer for realizing inverted sustainable planar PSCs in the future.
In contrast, Hu and co-authors [144] presented highly efficient perovskite solar cells with a poly(triarylamine) (PTAA) hole transport layer enabled by polymethyl methacrylate (PMMA) doping F4TCNQ. These PSCs exhibited a charger open-circuit voltage of 1.06 V, an FF of 76.3%, and a PCE of 17.9% for a concentration of the F4TCNQ solution of 0.025 mg/mL. In the same year, Liu and colleagues [59] described a way of doping the HTL with F4TCNQ and improving the initial stability of mesoporous spiro-OMeTAD PSCs. A highly stable mesoporous triple-cation PSC based on F4TCNQ with a very long T80 lifetime of more than 1 year (~380 days) for devices stored in air (RH ~ 40%) has been developed. This device exhibits a Voc of 1.05 V, a Jsc of 19.5 mA cm−2, an FF of 63.4%, and a PCE of 13.3%, which are higher than those for the undoped HTL (Voc, Jsc, FF, and PCE of 0.79 V, 18.9 mA cm−2, 43.4%, and 6.2%, respectively). Similar requests were observed in [145,146,147,148]: the conductivity in the spiro-OMeTAD layer is significantly influenced by the F4TCNQ doping level. Thus, the spiro-OMeTAD layer doped with F4TCNQ is primary responsible for the increased stability of the perovskite solar cells when compared to conventional hygroscopic dopants [149,150,151,152,153,154]. p-type doping of F4TCNQ hole-transporting materials enables the protection of the lower perovskite layer due to their hydrophobic properties (Figure 5).
Furthermore, new solar cells doped with TCNQ or its derivatives have an advantage in the construction of sustainable devices due to the stability of their photovoltaic performance under an air atmosphere. Thus, the optimization and engineering of TCNQ-doped perovskite solar cells could lead to highly efficient devices with long-term stability in the future.
TCNQ-based charge transfer materials also support carrier generation. As a result, an increase in the short-circuit current density related to conversion efficiency (η, %) is observed [155,156]. In 2023 [157], the fabrication and characterization of methylammonium perovskite solar cells using a TCNQ-doped decaphenylpentacyclosilane (DPPS)/CuPc(NH2)4 complex as a hole-transporting material were performed. According to Suzuki and co-authors, an increase in Jsc, FF, and Voc related to η was observed. This was caused by the charge transfer due to molecular interaction with p-orbital overlap between CuPc(NH2)4 and TCNQ molecules. This provided measured Voc, Jsc, FF, and η values of 0.766 V, 23.1 mA cm−2, 61.84%, and 10.9%, respectively. These DPPS/CuPc(NH2)4/TCNQ photovoltaic perovskite cells are characterized by stable photovoltaic performance with a good conversion efficiency even after 180 days compared to the above-described spiro-OMeTAD.
In summary, the use of TCNQ and its F4TCNQ derivative [4,20,21,22,23,24,25,57,58,59,60,100,114,115,116,137,141,144,145,156,157] is a promising doping technique as it allows for a non-hygroscopic additive. As a result, promising Voc, Jsc, FF, and PCE values for solar cells are being observed (Table 1). TCNQ-based OSCs, including perovskites, have extraordinary features such as a simple design, stability, flexibility, and a light weight compared to polymer solar cells. Through extensive work over the past several decades [64,75,77,79,86,87,88,92,93,94,95,131,132,133,134,135,138,139,140,146,147,148,149,150,151,152,153,154,155,158,159,160,161,162,163], the efficiency of solar cells, including PSCs doped with TCNQ, now rivals that of commercial photovoltaic materials.

3. TCNQ-Based Materials in Sustainable Green Energy Devices

Materials based on TCNQ are also suitable for use in portable and flexible green energy devices [167,168,169,170,171]. Starting from 2022 [172], multifunctional solar cells have seen significant improvements and now have an efficiency of 27.1% for flexible PSCs and of 33.9% for monolithic perovskite/Si solar cells [173]. The continuous research and development in this field has led to greater comprehension of the synthesis of new organic photovoltaic materials, resulting in more effective doping techniques. Modern strategies offer promising opportunities for achieving a higher PCE and advancing the field of TCNQ-derivative-based OSCs with enhanced efficiency and stability in the future. Thus, TCNQ and some TCNQ-based materials could contribute to the advancement of organic green energy PVs for the cost-effective realization of solar energy conversion.

4. Cathode TCNQ-Based Materials in Sustainable Batteries

Electrode materials [4,174,175,176,177,178] which use redox-active organic compounds as active components have attracted attention as promising alternatives to inorganic electrodes because they offer safety, stability, and environmental friendliness. Conductive TCNQ, as a small organic molecule with four cyan groups, is promoted today as a positive electrode material. The use of less than 20 wt.% conductive TCNQ can lead to high specific energy values [40,179,180,181].
Aqueous zinc ion batteries (AZIB) have been one of the most favorable power sources for portable and wearable electronic devices in recent years due to their low toxicity, excellent stability, and, most importantly, the appreciable specific energy (1086 Wh kg−1) and high theoretical capacity (820 mAh g−1) of zinc. As a result, metallic zinc anodes are regarded as one of the safest electrodes in aqueous batteries, while also having good compatibility with water and low redox potential (0.76 V vs. SHE) [182,183]. It is known that cyan groups can easily be reduced and regenerated via further oxidation [2,3,184,185]. For this reason, small-size organic electrode materials based on TCNQ [186] have also become a prominent choice for AZIBs. The stability of the cathode increases by confining TCNQ material inside the aqueous zinc batteries. Kundu et al. [187] reported similar small-size organic molecules as cathode materials for sustainable AZIBs. Chola and Nagarale [188] suggested the usefulness of TCNQ-based cathode materials for rechargeable aqueous zinc batteries with zinc plates as an anode. The cyclic voltammograms presented in [188] showed an oxidation peak at 1.4 V, whereas a reduction peak was observed at 0.65 V because of the coordination/decoordination of Zn2+ ions with TCNQ molecules. According to Chola and Nagarale, the stability and efficiency of a TCNQ-based cathode can be significantly increased by encapsulating the TCNQ inside a specially designed organic porous polymer, cyanuric chloride and pyrene (CCP). In contrast, TCNQ cathodes provide a higher capacity than a graphite positive electrode (ca. 60 mAh g−1).
Recently, TCNQ-based cathodes have also been used in AZIBs at room temperature. For example, Wang and his team [189] investigated highly purified TCNQ (p-TCNQ) as a cathode in AZIBs in a quite large temperature range of 25 °C to −40 °C. This p-TCNQ cathode delivered a high capacity of 160.7 mAh g−1 at 0 °C (a current density of 3 A g−1) and capacities of 172.9 and 96.6 mAh g−1 (a current density of 100 mA g−1) at −20 °C and −40 °C, respectively. This suggests better surface control with p-TCNQ and significantly better cycling stability than with TCNQ at various temperatures. In another work [190], Wang et al. described the charge transfer of TTF-TCNQ, TTF, and TCNQ organic sustainable electrode materials in AZIBs with an initial capacity of 153.4, 65.7, and 142.8 mAh g−1, respectively, at a current density of 1 A g−1.
A very strong TTF π-electron donor and its derivatives have recently emerged as key materials for new applications in electronics and molecular sensing due to their unusual electrical properties and smaller charge transfer resistance [45,98,100,190,191]. For example, according to Wang et al. [190], the TTF-TCNQ cathode material showed a far superior rate, cycling performance, and ionic diffusivity in ionic storage than each single-moiety counterpart, especially at low temperatures. It showed a specific capacity of 123.3 mAh g−1 at a current density of 3 A g−1 at −20 °C. Thus, Wang provides a new idea for the design of new organic electrode materials for sustainable high-performance batteries in the future.
In 2025, Ma and co-workers [192] demonstrated that because of TCNQ’s cyan groups, a high output voltage (1.15 V vs. Zn2+/Zn) with the incorporation of Na+ ions and an I/I3− redox mediator in a ZnSO4 aqueous electrolyte is possible. The TCNQ cathode demonstrated a significant improvement in rate capability and perfect cycle performance (2000 cycles with a decay rate of 0.0035% per cycle). The authors pointed to a higher reduction voltage with Zn2+ ions inserted into Zn(TCNQ)2. These results help us to understand the higher conductivity and ion storage capacity of TCNQ in the presence of Na+ and Zn2+ ions. Ref. [192] offered a promising approach for advancing sustainable energy storage systems. Also in 2025, Lee et al. [193] described the four-electron redox activity of an organic charge transfer complex (OCTC) TCNQ/PNZ (PNZ—phenazine) with superior cycle stability. It retained 88% of the maximum capacity of 122 mAh g−1 at a rate of 2 A g−1 over 100 cycles, using Zn-aqueous electrolytes (1 M Zn(TFSI)2 in H2O). Surprisingly, the full redox reaction achieves an unprecedentedly high electrode-level energy density, delivering ~10 mAh cm−2 of areal capacity (580 μm thick electrodes) in Zn batteries. Compared with electrodes of individual molecules, the TCNQ/PNZ OCTC could exhibit a significant enhancement in rate capability and cycle stability, which contributed to the electronic conductivity increasing by ~105 and ~20 times compared to PNZ and TCNQ molecules, respectively, and the suppressed dissolution in the electrolyte. When cycled within a broad voltage range of 1–3.8 V (vs. Li+/Li) encompassing redox activities of both entities (e.g., PNZ at 1.3, 1.9 V, and TCNQ at 2.6, 3.2 V (vs. Li+/Li), respectively), the electrochemical performance deteriorated rapidly, jeopardizing the expected benefits of the OCTC electrode in this trade-off. The OCT TCNQ/PNZ electrode exhibited extended redox capabilities to a four-electron transfer reaction with discharge voltages at 2.86, 2.92, 3.13, and 3.29 V (vs. Li+/Li) or 0.56, 0.62, 0.83, and 0.99 V (vs. Zn2+/Zn), respectively. The authors elucidate the complex interplay between organic electrodes and electrolytes in the charge storage mechanism. The same problem was explained in [194]. The results described in [193,194] highlight the importance of electrolyte design in developing sustainable organic electrode materials.
In the 1990s, lithium-ion batteries (LIBs), with their high energy efficiency of over 90%, were introduced to the battery market [195,196]. However, the main challenges related to their application in energy storage are their theoretical specific energy (400 Wh kg−1) and long charging time [197,198]. It is known that the ionic conductivity of non-aqueous electrolytes is two orders of magnitude lower than that of aqueous electrolytes [37,45,187,194,199,200,201]. In this regard, TCNQ, as one of the strongest π-acceptors [4,50,51,202,203,204,205], has also been proposed for the storage of Li ions. Fujihara and co-authors [206] developed a charge transfer TTF-TCNQ mixture as a cathode material without any additive. An overshooting potential was observed, especially at 0–10 mAh g–1 and 0.1–0.2 V. Similar behavior is often observed in cathode materials for LIBs [207,208].
It is known that the cycle stability of all organic conventional batteries is mainly governed by the aging and degradation of the electrolyte and electrodes due to repeated electrochemical reactions [209]. Copper-tetracyanoquinodimethane (CuTCNQ) is a perfect model of a sustainable electrode material [50,52,53,204] that can reversibly store Li+ ions under lower working voltages, leading to an increase in the life cycle length. CuTCNQ salts can exist in two polymorphic forms: the kinetically stable phase I, which has a needle morphology, and the low-conductivity thermodynamically stable phase II, which has a platelet morphology [210]. Both the Cu+ cation and the TCNQ- anion are electrochemically active, allowing CuTCNQ to participate in electron transfer redox reactions. For example, a self-supported CuTCNQ electrode for LIBs has been fabricated by Meng and co-authors [50]. The charge specific capacity of this electrode is 219.4 mAh g−1, reaching 280.9 mAh g−1 at a current density of 100 mA g−1 after 500 cycles, which was also obtained by Song et al. [211]. The reversible insertion and exiting of Li+ ions into the benzene rings of TCNQ ligands were observed. Moreover, the conductivity of the CuTCNQ film is 0.305 S cm−1 [50,212]. Thus, this CuTCNQ film can be used directly as a perfect candidate to design Li-based batteries in the future. For example, in [213], Yin and co-authors described a high-conductivity thin layer (TL) lithophilic large-surface-area CuTCNQ metal–organic framework (MOF). Recently, MOFs [214,215,216] emerged as a promising class of materials for battery applications due to the hybrid structure of their organic and inorganic parts. The CuTCNQ MOF TL has polar surfaces with a porous structure, providing it with a strong ability to adsorb Li+ ions from the electrolyte. This Li/CuTCNQ-MOF electrode has a very stable potential of around 13.1 mV, and therefore it could be a promising candidate in designing sustainable metal-based batteries.
In recent times, there has been a demand for advancements in the development of new positive electrode materials for reversible anion storage in dual-ion battery (DIB) cells. DIB technology using the appropriate electrolyte could be considered an environmentally friendly option for grid energy storage [217,218,219]. For example, Wang et al. [54] described CuTCNQ/lithium/sodium/graphite-based dual-ion batteries with ethylene carbonate/dimethyl carbonate/ethyl methyl carbonate as an electrolyte. The results of [54] demonstrate a higher working voltage, an excellent rate capacity of 76 mAh g−1 with a current density of 0.1 A g−1, and good reversibility of CuTCNQ DIB. Thus, TCNQ represents a promising type of sustainable material to overcome the limits of typical positive DIB electrodes. The CuTCNQ MOF positive electrode in DIB cells has also been investigated by Dühnen et al. [220]. Many known cathode materials for LIBs exhibit a conversion reaction with Li+ ions at high potentials. In contrast, the complete discharge of CuTCNQ (2.4 V vs. Li/Li+) results in a total specific capacity of 157 mAh g−1 in the first cycle [221,222,223]. According to Dühnen et al., a new CuTCNQ MOF represents an unexplored class of sustainable materials for anion storage in dual-ion cells.
In 2022, Geng et al. [55] presented some surprising results for F4TCNQ. A new solid-state high-voltage (3.4 V) Li/F4TCNQ battery was found to have a relatively high redox potential of 3.0 V vs. Li+/Li, which is higher than the redox potential of most organic cathode materials [224,225,226,227,228]!
Canals-Riclot and co-workers [229] designed an organic TCNQ all-solid-state lithium metal battery using a polymer covalent organic framework (COF)/polyethylene oxide (PEO) composite as a solid electrolyte. According to [229], the Li(TCNQ)-PEO-COF@LiI/PEO/Li cell was cycled at 70 °C and C/20 rate between 2.7 and 3.15 V vs. Li+/Li. A much better reversibility was observed, with 0.67 Li+ inserted per TCNQ unit at the second discharge (~88 mAh g−1) and 0.44 Li+ at the tenth discharge (~58 mAh g−1). The observed potentials for these two reversible charge/discharge processes are in good agreement with potentials reported by other authors [230] for the first and the second lithiation processes of TCNQ, respectively: TCNQ + Li+ + e ⟶ LiTCNQ and LiTCNQ + Li+ + e ⟶ Li2TCNQ (Figure 6). Therefore, very weak polarization was observed for these two processes (0.04 V and 0.1 V), representing a very good result for an all-solid-state cell. This improvement of the electrochemical behavior of the battery may be explained by the increased ionic conductivity of the composite polymer electrolyte containing COF@LiI.
Generally, highly porous π-electronic COFs can be used as materials for functional designs and as stable electrode materials at high potentials [230,231]. The electrochemical properties of COF-based materials have been significantly improved by introducing redox groups into their covalent organic framework. This favors the more widespread use of COFs in electrochemical sensors and other electronic devices [230,232,233,234]. TCNQ COFs also show a high electrical conductivity of 3.4 S m−1, which is a new record for 3D COF materials [235]. This presents a strong pretext for constructing new efficient energy storage devices based on TCNQ in the future.
Nowadays, because of lithium’s scarcity, batteries based on a sodium-ion (NIBs) [38,215,236,237,238] are also very promising from a practical point of view. NIB cells have huge potential in energy storage applications due to the abundance of the resources required for their creation, their low price, and their environmental friendliness. Organic compounds composed of C, H, and O are also cost-effective. Recently, CuTCNQ [202,204,205,239], as a promising high-capacity sustainable cathode material for both NIBs and potassium-ion batteries (KIBs), was investigated. In each charge/discharge process (voltage 2.0–4.1 V) a three-electron redox reaction occurs, which provides a very high reversible discharge capacity of 255 mAh g−1 [202]. As a result, cathode materials based on CuTCNQ have an extremely large capacity of 252 mAh g−1 and a specific energy density of more than 900 Wh kg−1. After oxidation, Na[Cu(TCNQ)2−] (CuII ⟶ CuI) forms a more conductive structure (phase I) of CuTCNQ with a record redox potential of 3.8 V vs. Na+/Na. According to [202], metal–organic CuTCNQ compounds present a new opportunity to achieve high energy density for NIBs and could provide a new way to design potential sustainable cathode materials for energy storage.
Recently, potassium ion batteries have drawn attention as a promising technology due to the relative abundance of potassium. A significant advantage of KIBs is the possibility of using graphite anodes in them, in contrast to NIBs. However, both KIBs and NIBs are easy to implement [240]. Potassium and sodium, which do not alloy with aluminum, can be used as anode current collectors, replacing copper in LIBs [241,242]. The abundant potassium resources (i.e., 1.5 wt. % in the Earth’s crust and 380.0 mg L−1 in seawater) mean that potassium-ion batteries can successfully be used in electric vehicles and in power grids for energy storage [243,244]. However, standard KIBs have low capacity even at voltages higher than 4.0 V. Therefore, modifying these devices by coating the electrode surfaces with TCNQ-based materials may enhance their performance. For example, the CuTCNQ organic material mentioned above shows a specific capacity of 244 mAh g−1, an average discharge potential of 2.8 V (vs. K+/K), and high performance of 70% of its initial capacity after 50 cycles [204].
Conductive TCNQ is a positive electrode material for rechargeable aluminum batteries as well. According to Guo et al. [245], TCNQ delivered a specific capacity of 180 mAh g−1 and a high discharging potential of 1.6 V (vs. Al3+/Al) for devices. The reversible capacity was still 115 mAh g−1 even after 2000 cycles at 500 mA g−1. There is thus an excellent opportunity to develop effective cyan-organic electrodes for sustainable aluminum batteries in the future. In contrast, Zhong et al. [43] found that the insertion/extraction of Al3+ into and out of TCNQ is highly reversible and leads to the improvement of the life cycle and specific capacity. The energy binding in the connection of TCNQ with Al3+ is −2.34 eV for the TCNQ0/TCNQ redox process. A TCNQ/PPy cathode exhibited a reversible capacity of 245.8 mAh g−1 at 0.5 A g−1 and a life cycle of 1000 rounds at 3 A g−1. This work demonstrates that adding highly valent cations in the presence of TCNQ is an effective strategy to enhance the electrochemical performance of organic electrodes in aqueous rechargeable batteries. Thus, aluminum TCNQ-based batteries [43,246,247,248] have also been regarded as the most effective for energy storage due to their high specific capacity. Inorganic positive electrode materials cannot properly meet the requirements of high-energy-density aluminum batteries [249,250,251]. TCNQ cyan groups [209,252,253] have also been used as the substituents for nucleophilic reactions, which allowed TCNQ-based organic molecules to coordinate with aluminum-chlorine coordination ions. TCNQ exhibits a great electron affinity, specific capacity (262 mAh g−1), and high reduction potential (1.96 V vs. Al3+/Al) [174,246]. Thus, TCNQ is the most appropriate cyan-organic material for sustainable aluminum batteries.
To summarize, organic TCNQ cathode materials are more attractive in view of their large surface area, good safety, and perfect physicochemical properties, including their low electrode potential, high redox capability, and high cycle stability (Scheme 3).
The features of TCNQ cathode materials for sustainable batteries are illustrated in Table 2.
As illustrated in the table above, TCNQ is successfully used as a positive electrode material in batteries. Nonetheless, it has high theoretic specific capacity and its performance can sometimes decline because of material dissolution [43].

5. TCNQ-Based Triboelectric Nanogenerators as Sustainable Wearable Power Sources for Energy Conversion

The increasing intelligence of wearable electronics has led to growing demand for higher energy levels but smaller volume or weight at the same time. In recent years [254,255,256,257,258,259,260], there has been a demand for triboelectric nanogenerators that can convert various types of mechanical energy into electrical energy, including body motions, water waves, or other vibrations. Through the contact electrification effect and the electrostatic induction effect, TENGs generate electrostatic charges and an external electric current, respectively [257,258,261,262]. Organic conductive materials can be used as electrification materials for the production of TENGs based on the tribovoltaic phenomenon [262,263,264]. Recently, several biodegradable and natural materials have been used to fabricate TENGs: cellulose, marine biomaterials, silk fibroin, petal roses, lotus, rice paper, etc. [31,265,266,267,268,269]. A wide range of carbon materials have also been tested for their suitability in energy harvesting devices. TCNQ conductive materials have been designed as sustainable electrode materials for TENGs due to their excellent electrical and physicochemical properties [270,271,272,273]. For example, Wang et al. [270] designed a TCNQ/PVA-based positive tribomaterial (PVA-polyvinyl alcohol) composed of Ni/Ag conductive tape as the top and bottom electrodes. This TCNQ-based TENG shows relatively high output power and can generate a Voc of 520 V and a Jsc of 218 mA m−2, regardless of the concentration of TCNQ. This demonstrates the versatile applications of TCNQ TENGs as sustainable capacitors for electronic devices and as wireless power transmitters. In another work, Zhang and co-workers [271] fabricated a new PVK:TCNQ(THF)/Ag (PVK—poly(9-vinylcarbazole) hybrid device with ideal 100% productivity. The TCNQ cyan groups coordinated very efficiently with Ag through the charge transfer between Ag and TCNQ: Agx + TCNQx + [Ag+TCNQ]n−x ↔ [Ag+TCNQ]n (0 < x ≤ n). According to [271], a new TCNQ-based hybrid device was found to have low threshold voltages of ~0.21 V and −0.3 V and a low operation power of 1.52 × 10−5 W, representing the lowest levels amongst organic devices. According to Zhang et al. [271], their device will inspire novel strategies for rationally designing multifunctional memory devices in the future.
A very strong F4TCNQ electron acceptor was used for the construction of a PEAI CABI/Spiro-PTAA-F4TCNQ-based TENG (PEAI—phenylethylammonium iodide; CABI—n-type lead-free Cu2AgBiI6 perovskite; PTAA—poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine]) by Choi et al. [274]. This nanogenerator shows a current density of 11.23 μA cm−2, a power density of 1.25 W m−2, and an upper voltage of 0.81 V. A p-type F4TCNQ dopant increases the charge mobility of the film. Jiao et al. [64] came to the same conclusion. This new F4TCNQ-based device is also temperature- and humidity-sensitive. The output voltage of the PEAI CABI/Spiro-PTAA-F4TCNQ-based TENG decreases with humidity. Thus, this TCNQ-based TENG can be used as a sustainable temperature and humidity sensor in the future.
All known examples of TCNQ-based nanogenerators meet the high sensitivity, safety, and biocompatibility requirements. Therefore, it is necessary to further develop highly efficient power management circuits and continue research that could lead to significant advances in the development and application of TCNQ-based TENGs. The future prospects of TCNQ-based TENGs are illustrated in Figure 7.

6. Conclusions and Perspectives

Since the beginning of the 21st century, due to their excellent physicochemical and electrical properties, TCNQ-based compounds have been exceptional candidates for various electronic applications. Over the last few decades, a wide range of TCNQ derivatives have been tested in this respect. This review presented the evolution of stable TCNQ-based materials for energy storage and conversion and the most important applications of TCNQ organic compounds as prospective sustainable cathode materials in solar cells, sustainable batteries, and TENGs.
This paper demonstrates that TCNQ organic compounds offer an interesting alternative to inorganic structures. TCNQ doping plays the dual role in improving the photoelectric properties and stability of perovskite devices due to the unusual redox behavior of TCNQ. CTCs based on TCNQ support charge transfer, yielding increases in short-circuit current density, which is related to conversion efficiency. These charge transfer processes are fundamental to creating advanced materials for application in electronics.
TCNQ-based electrode materials can be applied in a wide variety of energy storage devices such as non-aqueous Li+, Na+, K+-ion, dual-ion, multivalent-metal, and aqueous batteries. These electrodes can be fabricated with more than 90 wt.% of active material, enhancing the electrode-level specific energy. CuTCNQ is one of the best examples of a good electrode material, being able to reversibly store Li+/Na+/K+ ions under a lower working voltage. High-conductivity TCNQ positive electrode materials can also be used in rechargeable aluminum batteries. TCNQ-based batteries are regarded as the perfect sustainable candidates for low-cost and large-scale energy storage.
F4TCNQ, as a TCNQ derivative, is often used as an effective acceptor on electrode surfaces and a p-type dopant in sustainable energy devices due to its exceptionally high potential value of 5.2 eV.
Additionally, this review discusses the fundamental challenges involved in elevating TENGs to the next level, using TCNQ and some of its derivatives as sustainable materials. All TCNQ-based nanogenerators meet the high sensitivity, safety, and biocompatibility requirements.
In summary, using TCNQ-based materials with these unique properties expands the list of candidates for sustainable high-energy organic materials. This can be considered as a future alternative for high-performance solar cells, sustainable batteries, and high-sensitivity triboelectric nanogenerators. The presented data can be used as a guide for future research and should help to drive the development of new sustainable energy storage and conversion devices for a cleaner environment. This paper expands on a number of research papers discussing recent advances in energy storage and conversion. Hopefully, the information featured in this work will provide readers with a better understanding of TCNQ chemistry and illustrate the reasons why 7,7′,8,8′-tetracyanoquinodimethane can be considered as a sustainable material for sustainable energy device design.

Funding

This research is funded by the Jan Kochanowski University Rector Grant (Kielce, Poland) no. SUPB.RN.25.221.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

The author express their gratitude to Patrycja Rogala and Pawel Rodziewicz for their invaluable assistance in preparing this article.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AZIBaqueous zinc ion batteries
CABIn-type lead-free Cu2AgBiI6 perovskite
CCPcyanuric chloride and pyrene
COFcovalent organic framework
CTCcharge transfer complex
CVDchemical vapor deposition
DADQdiaminodicyanoquinodimethane
DIBdual-ion battery
DPPSdecaphenylpentacyclosilane
DSSCdye-sensitized solar cell
FFfill factor
HTLhole transport layer
ITOindium tin oxide
LIBlithium-ion battery
MOFmetal–organic framework
NIBsodium-ion battery
OCTCorganic charge transfer complex
OIHPorganic–inorganic halide perovskite
OSCorganic solar cell
PANIpolyaniline
Pcphthalocyanine
PCEpower conversion efficiency
PEAIphenylethylammonium iodide
PEOpolyethylene oxide
P3HTpoly(3-hexylthiophene)
PMMApolymethyl methacrylate
PNZphenazine
PPypolypyrrole
PSCperovskite solar cell
PThpolythiophene
PTAApoly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine]
PTTApoly(triarylamine)
PVApolyvinyl alcohol
PVKpoly(9-vinylcarbazole)
SHEstandard hydrogen electrode
STSCsolar thermal selective coating
TCNQ7,7′,8,8′-tetracyanoquinodimethane
TENGtriboelectric nanogenerator
THFtetrahydrofuran
TLthin layer
TTFtetrathiafulvalene

References

  1. Melby, L.R.; Harder, R.J.; Hertler, W.R.; Mahler, W.; Benson, R.E.; Mochel, W.E. Substituted quinodimethans. II. Anion-radical derivatives and complexes of 7,7,8,8-tetracyanoquinodimethan. J. Am. Chem. Soc. 1962, 84, 3374–3387. [Google Scholar] [CrossRef]
  2. Starodub, V.A.; Starodub, T.N. Radical anion salts and charge transfer complexes based on tetracyanoquinodimethane and other strong π-electron acceptors. Rus. Chem. Rev. 2014, 83, 391–438. [Google Scholar] [CrossRef]
  3. Starodub, W.; Starodub, T.; Chojnacki, J. Physicochemistry of Materials of Modern Electronics and Spintronics; Scientific Publishing House PWN SA: Warsaw, Poland, 2019; p. 360. ISBN 978-83-01-20734-2. (In Polish) [Google Scholar]
  4. Starodub, T.; Michalkiewicz, S. TCNQ and its Derivatives as Electrode Materials in Electrochemical Investigations—Achievement and Prospects: A Review. Materials 2024, 17, 5864. [Google Scholar] [CrossRef] [PubMed]
  5. Acker, D.S.; Hertler, W.R. Substituted quinodimethans. I. Preparation and chemistry of 7,7,8,8-tetracyanoquinodimethan. J. Am. Chem. Soc. 1962, 84, 3370–3374. [Google Scholar] [CrossRef]
  6. Ballester, L.; Gutierrez, A.; Perpinan, M.F.; Azcondo, M.T. Supramolecular architectures in low dimensional TCNQ compounds containing nickel and copper polyamine fragments. Coord. Chem. Rev. 1999, 190–192, 447–470. [Google Scholar] [CrossRef]
  7. Sánchez-Vergara, M.E.; Rios, C.; Jiménez-Sandoval, O.; Salcedo, R. A Comparative Study of the Semiconductor Behavior of Organic Thin Films: TCNQ-Doped Cobalt Phthalocyanine and Cobalt Octaethylporphyrin. Molecules 2020, 25, 5800. [Google Scholar] [CrossRef]
  8. O’Mullane, A.P.; Fay, N.; Nafady, A.; Bond, A.M. Preparation of metal-TCNQ charge-transfer complexes on conducting and insulating surfaces by photocrystallization. J. Am. Chem. Soc. 2007, 129, 2066–2073. [Google Scholar] [CrossRef] [PubMed]
  9. Starodub, T.N.; Barszcz, B.; Bednarski, W.; Mizera, A.; Starodub, V.A. Optical properties of RAS (N-CH3-2-NH2-5Cl-Py)(TCNQ)(CH3CN) solvate. J. Mol. Struct. 2020, 1201, 127121. [Google Scholar] [CrossRef]
  10. Tseng, T.C.; Urban, C.; Wang, Y.; Otero, R.; Tait, S.L.; Alcamí, M.; Ecija, D.; Trelka, M.; Gallego, J.M.; Lin, N.; et al. Charge-transfer-induced structural rearrangements at both sides of organic/metal interfaces. Nat. Chem. 2010, 2, 374–379. [Google Scholar] [CrossRef]
  11. Fernandez, C.A.; Martin, P.C.; Schaef, T.; Bowden, M.E.; Thallapally, P.K.; Dang, L.; Xu, W.; Chen, X.; McGrail, B.P. An electrically switchable metal-organic framework. Sci. Rep. 2014, 4, 6114–6120. [Google Scholar] [CrossRef]
  12. Tan, C.-H.; Ma, X.; Zhu, Q.-L.; Huang, Y.-H.; Wen, Y.-H.; Hu, S.-M.; Sheng, T.-L.; Wu, X.-T. Synthesis and characterization of cobalt(III) cyanide complexes: Cobalt participation in the decomposition of radical anion of TCNQ. CrystEngComm 2012, 14, 8708–8713. [Google Scholar] [CrossRef]
  13. Kaim, W.; Moscherosch, M. The coordination chemistry of TCNE, TCNQ and related polynitrile π acceptors. Coord. Chem. Rev. 1994, 129, 157–193. [Google Scholar] [CrossRef]
  14. Starodub, T.; Starodub, V. Sole anionorodnikowe TCNQ z kationami kompleksowymi metali przejściowych. Wiad. Chem. 2020, 74, 761–796. [Google Scholar]
  15. Ungor, O.; Shatruk, M. Transition metal complexes with fractionally charges TCNQ radical anions as structural templates for multifunctional molecular conductors. Polyhedron 2020, 177, 114254–114271. [Google Scholar] [CrossRef]
  16. Mao, G.; Kilani, M.; Ahmed, M. Review—Micro/Nanoelectrodes and Their Use in Electrocrystallization: Historical Perspective and current trend. J. Electrochem. Soc. 2022, 169, 022505–022517. [Google Scholar] [CrossRef]
  17. Saito, G.; Yoshida, Y. Frontiers of Organic Conductors and Superconductors. Top. Curr. Chem. 2012, 312, 67–126. [Google Scholar] [CrossRef]
  18. Jerome, D. Organic conductors: From charge density wave TTF-TCNQ to superconducting (TMTSF)2PF6. Chem. Rev. 2004, 104, 5565–5591. [Google Scholar] [CrossRef]
  19. Horiuchi, S.; Tokura, Y. Organic ferroelectrics. Nat. Mater. 2008, 7, 357–366. [Google Scholar] [CrossRef]
  20. Li, Q.-S.; Ye, B.-C.; Liu, B.-X.; Zhong, J.-J. Inprovement of the performance of H2O2 oxidation at low working potential by incorporating TTF-TCNQ into a platinum wire electrode for glucose determination. Biosens. Bioelectron. 1999, 14, 327–334. [Google Scholar] [CrossRef]
  21. Palmisano, F.; Zambonin, P.; Centonze, D.; Quinto, M. A disposable, reagentless, third-generation glucose biosensor based on overoxidized poly(pyrrole)/tetrathiafulvalene-tetracyanoquinodimethane composite. Anal. Chem. 2002, 74, 5913–5918. [Google Scholar] [CrossRef]
  22. Hernández-Cruz, M.; Galán-Vidal, C.A.; Álvarez-Romero, G.A.; Ramírez-Silva, M.T.; Páez-Hernández, M.E.; González-Vidal, J.L. Behavior of Two and Three Electrode Configuration and Different Mediators in Working Electrode on Development of Disposable Screen-Printing Biosensors for Determination of Free Cholesterol. J. Mex. Chem. Soc. 2013, 57, 47–53. [Google Scholar] [CrossRef]
  23. Yuge, R.; Yudasaka, M.; Maigne, A.; Tomonari, M.; Miyawaki, J.; Kubo, Y.; Imai, H.; Ichihashi, T.; Iijima, S. Adsorption Phenomena of Tetracyano-p-quinodimethane on Single-Wall Carbon Nanohorns. J. Phys. Chem. C 2008, 112, 5416–5422. [Google Scholar] [CrossRef]
  24. Khanra, P.; Lee, C.-N.; Kuila, T.; Kim, N.H.; Park, M.J.; Lee, J.H. 7,7,8,8-tetracyanoquinodimethane-assisted one-step electrochemical exfoliation of graphite and its performance as an electrode material. Nanoscale 2014, 6, 4864–4873. [Google Scholar] [CrossRef]
  25. Yuan, B.; Xu, C.; Zhang, R.; Lv, D.; Li, S.; Zhang, D.; Liu, L.; Fernandez, C. Glassy carbon electrode modified with 7,7,8,8-tetracyanoquinodimethane and graphene oxide triggered a synergistic effect: Low-potential amperometric detection of reduced glutathione. Biosens. Bioelectron. 2017, 96, 1–7. [Google Scholar] [CrossRef]
  26. Sivasankaran, R.P.; Das, P.K.; Arunachalam, M.; Kanase, R.S.; Il Park, Y.; Seo, J.; Kang, S.H. TiO2 Nanotube Arrays Decorated with Reduced Graphene Oxide and Cu–Tetracyanoquinodimethane as Anode Materials for Photoelectrochemical Water Oxidation. ACS Appl. Nano Mater. 2021, 4, 13218–13233. [Google Scholar] [CrossRef]
  27. Shimomura, S.; Kitagawa, S. Soft porous crystal meets TCNQ: Charge transfer-type porous coordination polymers. J. Mater. Chem. 2011, 21, 5537–5546. [Google Scholar] [CrossRef]
  28. Miyasaka, H.; Izawa, T.; Takahashi, N.; Yamashita, M.; Dunbar, K.R. Long-Range Ordered Magnet of a Charge-Transfer Ru24+/TCNQ Two-Dimensional Network Compound. J. Am. Chem. Soc. 2006, 128, 11358–11359. [Google Scholar] [CrossRef]
  29. Miyasaka, H.; Motokawa, N.; Matsunaga, S.; Yamashita, M.; Sugimoto, K.; Mori, T.; Toyota, N.; Dunbar, K.R. Control of Charge Transfer in a Series of Ru2II,II/TCNQ Two-Dimensional Networks by Tuning the Electron Affinity of TCNQ Units: A Route to Synergistic Magnetic/Conducting Materials. J. Am. Chem. Soc. 2010, 132, 1532–1544. [Google Scholar] [CrossRef]
  30. Li, Q.; Wang, Y.; Yan, P.; Hou, G.; Li, G. Two 7,7,8,8-tetracyanoquinodimethane lead and zinc complexes featuring 3D and 0D structure: Synthesis, structure and electrochemical properties. Inorg. Chim. Acta 2014, 413, 32–37. [Google Scholar] [CrossRef]
  31. Li, Y.; Liu, X.; Ren, Z.; Luo, J.; Zhang, C.; Cao, C.; Yuan, H.; Pang, Y. Marine biomaterial-based triboelectric nanogenerators: Insights and applications. Nano Energy 2024, 119, 109046–109068. [Google Scholar] [CrossRef]
  32. Sarma, S.J.; Brar, S.K.; Sydney, E.B.; Le Bihan, Y.; Buelna, G.; Soccol, C.R. Microbial hydrogen production by bioconversion of crude glycerol: A review. Int. J. Hydrogen Energy 2012, 37, 6473–6490. [Google Scholar] [CrossRef]
  33. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The role of renewable energy in the global energy transformation. Energy Strategy Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  34. Lu, Y.; Chen, J. Prospects of organic electrode materials for practical lithium batteries. Nat. Rev. Chem. 2020, 4, 127–142. [Google Scholar] [CrossRef] [PubMed]
  35. Shea, J.; Luo, C. Organic Electrode Materials for Metal Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 5361–5380. [Google Scholar] [CrossRef] [PubMed]
  36. Xie, J.; Zhang, Q. Recent Progress in Multivalent Metal (Mg, Zn, Ca, and Al) and Metal-Ion Rechargeable Batteries with Organic Materials as Promising Electrodes. Small 2019, 15, 1805061–1805080. [Google Scholar] [CrossRef]
  37. Huang, J.; Dong, X.; Guo, Z.; Wang, Y. Progress of Organic Electrodesing Aqueous Electrolyte for Energy Storage and Conversion. Angew. Chem. Int. Ed. 2020, 59, 18322–18333. [Google Scholar] [CrossRef]
  38. Gannett, C.N.; Kim, J.; Tirtariyadi, D.; Milner, P.J.; Abruna, H.D. Investigation of ion-electrode interactions of linear polyimides and alkali metal ions for next generation alternative-ion batteries. Chem. Sci. 2022, 13, 9191–9201. [Google Scholar] [CrossRef]
  39. Cha, H.; Kim, H.N.; An, T.K.; Kang, M.S.; Kwon, S.K.; Kim, Y.H.; Park, C.E. Effects of Cyano-Substituents on the Molecular Packing Structures of Conjugated Polymers for Bulk-Heterojunction Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 15774–15782. [Google Scholar] [CrossRef]
  40. Häupler, B.; Burges, R.; Janoschka, T.; Jähnert, T.; Wild, A.; Schubert, U.S. PolyTCNQ in organic batteries: Enhanced capacity at constant cell potential using two-electron-redox-reactions. J. Mater. Chem. A 2014, 2, 8999–9001. [Google Scholar] [CrossRef]
  41. Neelamraju, B.; Watts, K.E.; Pemberton, J.E.; Ratcliff, E.L. Correlation of Coexistent Charge Transfer States in F4TCNQ-Doped P3HT with Microstructure. J. Phys. Chem. Lett. 2018, 9, 6871–6877. [Google Scholar] [CrossRef]
  42. Bavdane, P.P.; Dave, V.; Sreenath, S.; Madiyan, P.; Nagarale, R.K. Redox-active organic molecule encapsulated MWCNT catholyte for aqueous zinc flow battery. Next Energy 2025, 9, 100379. [Google Scholar] [CrossRef]
  43. Zhong, Y.; Li, Y.; Meng, J.; Lin, X.; Huang, Z.; Shen, Y.; Huang, Y. Boosting the cyclability of tetracyanoquinodimethane (TCNQ) as cathode material in aqueous battery with high valent cation. Energy Storage Mater. 2021, 43, 492–498. [Google Scholar] [CrossRef]
  44. Meng, J.; Tang, Q.; Zhou, L.; Zhao, C.; Chen, M.; Shen, Y.; Zhou, J.; Feng, G.; Shen, Y.; Huang, Y. A Stirred Self-Stratified Battery for Large-Scale Energy Storage. Joule 2020, 4, 953–966. [Google Scholar] [CrossRef]
  45. Wang, S.; Hu, N.; Huang, Y.; Deng, W. Charge-transfer complex promoted energy storage performance of single-moiety organic electrode materials in aqueous zinc-ion battery at low temperatures. Appl. Surf. Sci. 2023, 619, 156725–156733. [Google Scholar] [CrossRef]
  46. Yang, M.; Leon, N.; Pan, B.; Yu, Z.; Cheng, L.; Liao, C. Mechanistic insights in quinone-based zinc batteries with nonaqueous electrolytes. J. Electrochem. Soc. 2020, 167, 100536. [Google Scholar] [CrossRef]
  47. Hu, Y.; Yu, Q.; Tang, W.; Cheng, M.; Wang, X.; Liu, S.; Gao, J.; Wang, M.; Xiong, M.; Hu, J.; et al. Ultra-Stable, Ultra-Long-Lifespan and Ultra-High-Rate Na-Ion Batteries Using Small-Molecule Organic Cathodes. Energy Storage Mater. 2021, 41, 738–747. [Google Scholar] [CrossRef]
  48. Mike, J.F.; Lutkenhaus, J.L. Recent advances in conjugated polymer energy storage. J. Polym. Sci. B Polym. Phys. 2013, 51, 468–480. [Google Scholar] [CrossRef]
  49. Xie, J.; Gu, P.; Zhang, Q. Nanostructured Conjugated Polymers: Toward High-Performance Organic Electrodes for Rechargeable Batteries. ACS Energy Lett. 2017, 2, 1985–1996. [Google Scholar] [CrossRef]
  50. Meng, C.; Chen, T.; Fang, C.; Huang, Y.; Hu, P.; Tong, Y.; Bian, T.; Zhang, J.; Wang, Z.; Yuan, A. Multiple Active Sites: Lithium Storage Mechanism of Cu-TCNQ as an Anode Material for Lithium-Ion Batteries. Chem. Asian J. 2019, 14, 4289–4295. [Google Scholar] [CrossRef]
  51. Li, B.; Zhao, J.; Zhang, Z.; Zhao, C.; Sun, P.; Bai, P.; Yang, J.; Zhou, Z.; Xu, Y. Electrolyte regulated solid-electrolyte interphase enables long cycle life performance in organic cathodes for potassium-ion batteries. Adv. Funct. Mater. 2019, 29, 1807137. [Google Scholar] [CrossRef]
  52. Huang, Y.; Fang, C.; Zeng, R.; Liu, Y.; Zhang, W.; Wang, Y.; Liu, Q.; Huang, Y. In Situ-Formed Hierarchical Metal–Organic Flexible Cathode for High-Energy Sodium-Ion Batteries. ChemSusChem 2017, 10, 4704–4708. [Google Scholar] [CrossRef]
  53. Fang, C.; Ye, Z.; Wang, Y.; Zhao, X.; Huang, Y.; Zhao, R.; Liu, J.; Han, J.; Huang, Y. Immobilizing an organic electrode material through p-p interaction for high performance Li-organic batteries. J. Mater. Chem. A 2019, 7, 22398–22404. [Google Scholar] [CrossRef]
  54. Wang, H.; Wu, Q.; Wang, Y.; Lv, X.; Wang, H.-G. A redox-active metal–organic compound for lithium/sodium-base dual-ion batteries. J. Colloid Interface Sci. 2022, 606, 1024–1030. [Google Scholar] [CrossRef]
  55. Geng, Z.; Shi, W.; Wang, J.; Zhang, X.; Hu, N.; Wang, S.; Deng, W. A high-voltage solid-state organic lithium battery based on 2,3,5,6-tetrafluoro-7,7,8,8-tetracyanoquinodimethane cathode. Int. J. Energy Res. 2022, 46, 12316–12322. [Google Scholar] [CrossRef]
  56. Bitenc, J.; Pirnat, K.; Luzanin, O.; Dominko, R. Organic Cathodes, a Path toward Future Sustainable Batteries: Mirage or realistic Future? Chem. Mater. 2024, 36, 1025–1040. [Google Scholar] [CrossRef] [PubMed]
  57. Abicho, S.; Hailegnaw, B.; Mayr, F.; Cobet, M.; Yumusak, C.; Sergawi, A.; Yohannes, T.; Kaltenbrunner, M.; Scharber, M.C.; Workneh, G.A. The role of TCNQ for surface and interface passivation in inverted perovskite solar cells. Mater. Ren. Sustain. Energy 2025, 14, 11. [Google Scholar] [CrossRef]
  58. Schloemer, T.H.; Christians, J.A.; Luther, J.M.; Sellinger, A. Doping strategies for small molecule organic hole-transport materials: Impacts on perovskite solar cell performance and stability. Chem. Sci. 2019, 10, 1904–1935. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, M.; Dahlström, S.; Ahläng, C.; Wilken, S.; Degterev, A.; Matuhina, A.; Hadadian, M.; Markkanen, M.; Aitola, K.; Kamppinen, A.; et al. Beyond hydrophobicity: How F4-TCNQ doping of the hole transport material improves stability of mesoporous triple-cation perovskite solar cells. J. Mater. Chem. A 2022, 10, 11721–11731. [Google Scholar] [CrossRef]
  60. Mohitkar, A.; Renuka, H.; Goel, S.; Jayanty, S. Efficient Standalone Flexible Small Molecule Organic Solar Cell Devices: Structure-Performance Relation Among Tetracyanoquinodimethane Derivatives. ACS Omega 2023, 8, 40836–40847. [Google Scholar] [CrossRef]
  61. Mi, B.X.; Gao, Z.Q.; Cheah, K.W.; Chen, C.H. Organic Light-Emitting Diodes Using 3,6-Difluoro-2,5,7,7,8,8-Hexacyanoquinodimethane as p-Type Dopant. Appl. Phys. Lett. 2009, 94, 073507–073510. [Google Scholar] [CrossRef]
  62. Song, T.; Chen, Q.; Zhou, H.; Jiang, C.; Wang, H.; Yang, Y.M.; Liu, Y.; You, J.; Yang, Y. Perovskite solar cells: Film formation and properties. J. Mater. Chem. A 2015, 3, 9032–9050. [Google Scholar] [CrossRef]
  63. Jung, H.S.; Park, N. Perovskite Solar Cells: From Materials to Devices. Small 2015, 11, 10–25. [Google Scholar] [CrossRef] [PubMed]
  64. Jiao, Y.; Lin, Z.; Xu, Y.; Zhang, B.; Guo, X.; Hu, Z.; Zhao, X.; Hao, Y.; Chang, J. Dynamic p-n junction direct current-generating triboelectric nanogenerators based on lead-free perovskite. Nano Energy 2025, 138, 110857. [Google Scholar] [CrossRef]
  65. Li, Q.X.; Shi, C.; Huang, M.L.; Wei, X.; Yan, H.; Yang, C.L.; Yuan, A.H. B- and N-embedded color-tunable phosphorescent iridium complexes and B-N Lewis adducts with intriguing structural and optical changes. Chem. Sci. 2019, 10, 3257–3263. [Google Scholar] [CrossRef] [PubMed]
  66. Zhou, Z.K.; Zhou, J.P.; Gai, L.Z.; Yuan, A.H.; Shen, Z. Low-cost CuNi@MIL-101 as an excellent catalyst toward cascade reaction: Integration of ammonia borane dehydrogenation with nitroarene hydrogenation. Chem. Commun. 2017, 53, 6621–6624. [Google Scholar] [CrossRef]
  67. Njoh, A.J.; Etta, S.; Essia, U.; Ngyah-Etchutambe, I.; Enomah, L.E.; Tabrey, H.T.; Tarke, O.M. Implications of institutional frameworks for renewable energy policy administration: Case study of the Esaghem, Cameroon community PV solar electrification project. Energy Pol. 2019, 128, 17–24. [Google Scholar] [CrossRef]
  68. Talaat, M.; Farahat, M.A.; Elkholy, M.H. Renewable power integration: Experimental and simulation study to investigate the ability of integrating wave, solar and wind energies. Energy 2019, 170, 668–682. [Google Scholar] [CrossRef]
  69. Ajayan, J.; Nirmal, D.; Mohankumar, P.; Saravanan, M.; Jagadesh, M.; Arivazhagan, L. A review of photovoltaic performance of organic/inorganic solar cells for future renewable and sustainable energy technologies. Superlattices Microstruct. 2020, 143, 106549–106602. [Google Scholar] [CrossRef]
  70. Alferov, Z.I.; Andreev, V.M.; Kagan, M.B.; Protasov, I.I.; Trofim, V.G. Solar-energy converters based on p-n AlxGa12xAs-GaAs heterojunctions. Fiz. Tekh. Poluprovodn. 1970, 4, 2378–2379. [Google Scholar]
  71. Hawkins, H.M. Concepts of Digital Electronics; Tab Books: Blue Ridge Summit, PA, USA, 1983; p. 231. ISBN 0830615318. [Google Scholar]
  72. Merritt, V.Y. Organic Photovoltaic Materials: Squarylium and Cyanine-TCNQ Dyes. IBM J. Res. Dev. 1978, 22, 353–371. [Google Scholar] [CrossRef]
  73. Weber, D. CH3NH3SnBrxI3-x (x = 0-3), ein Sn(II)-System mit kubischer Perowskitstruktur/CH3NH3SnBrxI3-x(x = 0-3), a Sn(II)-System with Cubic Perovskite Structure. Z. Naturforschung 1978, 33, 862–865. [Google Scholar] [CrossRef]
  74. Zweibel, K. Thin Films: Past, Present and Future; NREL/TP-413-7486; National Renewable Energy Laboratory: Golden, CO, USA, 1995. [CrossRef]
  75. Zhang, Y.; Elawad, M.; Yu, Z.; Jiang, X.; Lai, J.; Sun, L. Enhanced performance of perovskite solar cells with P3HT hole-transporting materials via molecular p-type doping. RSC Adv. 2016, 110, 108888–108895. [Google Scholar] [CrossRef]
  76. Cao, H.; He, W.; Mao, Y.; Lin, X.; Ishikawa, K.; Dickerson, J.H.; Hess, W.P. Recent progress in degradation and stabilization of organic solar cell. J. Power Sources 2014, 264, 168–183. [Google Scholar] [CrossRef]
  77. Zayed, M.E.; Elsheikh, A.H.; Essa, F.A.; Elbanna, A.M.; Li, W.; Zhao, J. High-temperature solar selective absorbing coatings for concentrated solar power systems. Sustain. Mater. Green Proc. Energy Conv. 2022, 12, 361–398. [Google Scholar] [CrossRef]
  78. Li, Y.; Xu, G.; Cui, C.; Li, Y. Flexible and Semitransparent Organic Solar Cells. Adv. Energy Mater. 2017, 8, 1701791–1701819. [Google Scholar] [CrossRef]
  79. Yan, L.L.; Han, C.; Shi, B.; Zhao, Y.; Zhang, X.D. A review on the crystalline silicon bottom cell for monolithic perovskite/silicon tandem solar cells. Mater. Today Nano 2019, 7, 100045–100058. [Google Scholar] [CrossRef]
  80. Cho, D.H.; Jo, H.S.; Lee, W.J.; Kim, T.G.; Shin, B.; Yoon, S.S.; Chung, Y.D. Enhanced electrical conductivity of transparent electrode using metal microfiber networks for griddles thin-film solar cells. Sol. Energy Mater. Sol. Cells 2019, 200, 109998. [Google Scholar] [CrossRef]
  81. Gong, J.; Sumathy, K.; Qiao, Q.; Zhou, Z. Review on dye-sensitized solar cells (DSSCs): Advanced techniques and research trends. Renew. Sustain. Energy Rev. 2017, 68, 234–246. [Google Scholar] [CrossRef]
  82. Sharma, S.; Siwach, B.; Ghoshal, S.K.; Mohan, D. Dye sensitized solar cells: From genesis to recent drifts. Renew. Sustain. Energy Rev. 2017, 70, 529–537. [Google Scholar] [CrossRef]
  83. Ikpesu, J.E.; Iyuke, S.E.; Daramola, M.; Okewale, A.O. Synthesis of improved dye-sensitized solar cell for renewable energy power generation. Sol. Energy 2020, 206, 918–934. [Google Scholar] [CrossRef]
  84. Sun, Z.Z.; Feng, S.; Ding, W.L. Integrated probing the influence of dye acceptor with several electron withdrawing groups for dye-sensitized solar cells. Sol. Energy 2020, 195, 491–498. [Google Scholar] [CrossRef]
  85. Heng, P.; Mao, L.; Guo, X.; Wang, L.; Zhang, J. Accurate estimation of the photoelectric conversion efficiency of a series of anthracene-based organic dyes for dye-sensitized solar cells. J. Mater. Chem. 2020, 8, 2388–2399. [Google Scholar] [CrossRef]
  86. Zayed, M.E. Recent Advances in Solar Thermal Selective Coatings for Solar power Applications: Technology Categorization, Preparation methods, and Induced aging mechanisms. Appl. Sci. 2024, 14, 8438. [Google Scholar] [CrossRef]
  87. Jena, A.K.; Kulkarni, A.; Miyasaka, T. Halide Perovskite Photovoltaics: Background, Status, and Future Prospects. Chem. Rev. 2019, 119, 3036–3103. [Google Scholar] [CrossRef]
  88. Li, X.; Yu, H.; Liu, Z.; Huang, J.; Ma, X.; Liu, Y.; Sun, Q.; Dai, L.; Ahmad, S.; Shen, Y.; et al. Progress and Challenges Toward Effective Flexible Perovskite Solar Cells. Nano-Micro Lett. 2023, 15, 206. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, C.R.; Zhang, Y.; Li, X.Y.; Wang, W.; Gong, J.J.; Liu, Z.J.; Chen, H.S. The bis-dimethylfluoreneaniline organic dye sensitizers for solar cells: A theoretical study and design. J. Mol. Graph. 2019, 88, 23–31. [Google Scholar] [CrossRef]
  90. Walki, S.; Naik, L.; Savanur, H.M.; Yogananda, K.C.; Naik, S.; Ravindra, M.K.; Malimat, G.H.; Mahadevan, K.M. Design of new imidazole-derivative dye having donor-π-acceptor moieties for highly efficient organic-dye-sensitized solar cells. Optik 2019, 208, 164074. [Google Scholar] [CrossRef]
  91. Wang, L.; Zhang, J.; Duan, Y.C.; Pan, Q.Q.; Wu, Y.; Geng, Y.; Su, Z.M. Theoretical insights on the rigidified dithiophene effects on the performance of near-infrared cis-squaraine-based dye-sensitized solar cells with panchromatic absorption. J. Photochem. Photobiol. A 2019, 369, 150–158. [Google Scholar] [CrossRef]
  92. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.K.; Moon, C.S.; Jeon, N.J.; Baena, J.P.C.; et al. Efficient perovskite solar cells via improved carrier management. Nature 2021, 590, 587–593. [Google Scholar] [CrossRef] [PubMed]
  93. Dong, Q.; Zhu, C.; Chen, M.; Jiang, C.; Guo, J.; Feng, Y.; Dai, Z.; Yadavalli, S.K.; Hu, M.; Cao, X.; et al. Interpenetrating interfaces for efficient perovskite solar cells with high operational stability and mechanical robustness. Nat. Commun. 2021, 12, 973. [Google Scholar] [CrossRef]
  94. Sharma, R.; Sharma, A.; Agarwal, S.; Dhaka, M.S. Stability and efficiency issues, solutions and advancements in perovskite solar cells: A review. Sol. Energy 2022, 244, 516–535. [Google Scholar] [CrossRef]
  95. Chen, J.; Yang, Y.; Dong, H.; Li, J.; Zhu, X.; Xu, J.; Pan, F.; Yuan, F.; Dai, J.; Jiao, B.; et al. Highly efficient and stable perovskite solar cells enabled by lowdimensional perovskitoids. Sci. Adv. 2022, 8, eabk2722. [Google Scholar] [CrossRef] [PubMed]
  96. Williams, G.; Sutty, S.; Aziz, H. Interplay between efficiency and device architecture for small molecule organic solar cells. Phys. Chem. Chem. Phys. 2014, 16, 11398–11408. [Google Scholar] [CrossRef]
  97. Vohra, V.; Kawashima, K.; Kakara, T.; Koganezawa, T.; Osaka, I.; Takimiya, K.; Murata, H. Efficient inverted polymer solar cells employing favourable molecular orientation. Nat. Photonics 2015, 9, 403–408. [Google Scholar] [CrossRef]
  98. Alves, H.; Molinari, A.S.; Xie, H.; Alberto, F.M. Metallic Conduction at Organic Charge-Transfer Interfaces. Nat. Mater. 2008, 7, 574–580. [Google Scholar] [CrossRef]
  99. Kirtley, J.R.; Mannhart, J. Organic electronics: When TTF met TCNQ. Nat. Mater. 2008, 7, 520–521. [Google Scholar] [CrossRef]
  100. Suzuki, A.; Ohtsuki, T.; Oku, T.; Akiyama, T. Fabrication and characterization of tetracyanoquinodimethane/phthalocyanine solar cells. Mater. Sci. Engin. B 2012, 177, 877–881. [Google Scholar] [CrossRef]
  101. Hsu, C.-L.; Lin, C.-T.; Huang, J.-H.; Chu, C.-W.; Wei, K.-H.; Li, L.-J. Layer-by-Layer Graphene/TCNQ Stacked Films as Conducting Anodes for Organic Solar Cells. Asc Nano 2012, 6, 5031–5039. [Google Scholar] [CrossRef]
  102. Krebs, F.C. All Solution Roll-to-Roll Processed Polymer Solar Cells Free from Indium-Tin-Oxide and Vacuum Coating Steps. Org. Electron. 2009, 10, 761–768. [Google Scholar] [CrossRef]
  103. Liu, C.; Li, Z.; Zhang, Z.; Zhang, X.; Shen, L.; Guo, W.; Zhang, L.; Long, Y.; Ruan, S. Improving the charge carrier transport of organic solar cells by incorporating a deep energy level molecule. Phys. Chem. Chem. Phys. 2017, 19, 245–250. [Google Scholar] [CrossRef] [PubMed]
  104. Yim, K.H.; Whiting, G.L.; Murphy, C.E.; Halls, J.J.M.; Burroughes, J.H.; Friend, R.H.; Kim, J.S. Controlling Electrical properties of Conjugated polymers via a Solution-Based p-Type Doping. Adv. Mater. 2008, 20, 3319–3324. [Google Scholar] [CrossRef]
  105. Hu, P.; Li, H.; Li, Y.; Jiang, H.; Kloc, C. Single-crystal growth, structures, charge transfer and transport properties of anthracene-F4TCNQ and tetracene-F4TCNQ charge-transfer compounds. CrystEngComm 2017, 19, 618–624. [Google Scholar] [CrossRef]
  106. Untilova, V.; Zeng, H.; Durand, P.; Herrmann, L.; Leclerc, N.; Brinkmann, M. Intercalation and Ordering of F6TCNNQ and F4TCNQ Dopants in Regioregular Poly(3-hexylthiophene) Crystals: Impact on Anisotropic Thermoelectric Properties of Oriented Thin Films. Macromolecules 2021, 54, 6073–6084. [Google Scholar] [CrossRef]
  107. Gao, W.; Kahn, A. Electronic structure and current injection in zinc phthalocyanine doped with tetrafluorotetracyanoquinodimethane: Interface versus bulk effects. Org. Electron. 2002, 3, 53–63. [Google Scholar] [CrossRef]
  108. Su, Y.H.; Ke, Y.F.; Cai, S.L.; Yao, Q.Y. Surface plasmon resonance of layer-by-layer gold nanoparticles induced photoelectric current in environmentally-friendly plasmon-sensitized solar cell. Light Sci. Appl. 2012, 1, e14–e19. [Google Scholar] [CrossRef]
  109. Chen, X.; Jia, B.H.; Zhang, Y.A.; Gu, M. Exceeding the limit of plasmonic light trapping in textured screen-printed solar cells using Al nanoparticles and wrinkle-like graphene sheets. Light Sci. Appl. 2013, 2, e92–e98. [Google Scholar] [CrossRef]
  110. Holman, Z.C.; Wolf, S.D.; Ballif, C. Improving metal reflectors by suppressing surface plasmon polaritons: A priori calculation of the internal reflectance of a solar cell. Light Sci. Appl. 2013, 2, e106–e122. [Google Scholar] [CrossRef]
  111. Chandaluri, C.G.; Radhakrishnan, T.P. Zwitterionic Diaminodicyanoquinodimethanes with Enhanced Blue-Green Emission in the Solid State. Opt. Mater. 2011, 34, 119–125. [Google Scholar] [CrossRef]
  112. Boyineni, A.; Jayanty, S. Supramolecular Helical Self-Assemblies and Large Stokes Shift in 1-(2-Cyanophenyl)Piperazine and 4-Piperidinopiperidine Bis-Substituted Tetracyanoquinodimethane Fluorophores. Dye. Pigm. 2014, 101, 303–311. [Google Scholar] [CrossRef]
  113. Radhakrishnan, T.P. Molecular Structure, Symmetry, and Shape as Design Elements in the Fabrication of Molecular Crystals for Second Harmonic Generation and the Role of Molecules-in-Materials. Acc. Chem. Res. 2008, 41, 367–376. [Google Scholar] [CrossRef]
  114. Mohammadtaheri, M.; Ramanathan, R.; Bansal, V. Emerging Applications of Metal-TCNQ Based Organic Semiconductor Charge Transfer Complexes for Catalysis. Catal. Today 2016, 278, 319–329. [Google Scholar] [CrossRef]
  115. Zhang, Z.; Gou, G.; Wan, J.; Li, H.; Wang, M.; Li, L. Synthesis, Structure, and Significant Energy Gap Modulation of Symmetrical Silafluorene-Cored Tetracyanobutadiene and Tetracyanoquinodimethane Derivatives. J. Org. Chem. 2022, 87, 2470–2479. [Google Scholar] [CrossRef]
  116. Matsuoka, S.; Ogawa, K.; Ono, R.; Nikaido, K.; Inoue, S.; Higashino, T.; Tanaka, M.; Tsutsumi, J.; Kondo, R.; Kumai, R.; et al. Highly Stable and Isomorphic Donor-Acceptor Stacking in a Family of n-Type Organic Semiconductors of BTBT-TCNQ Derivatives. J. Mater. Chem. C 2022, 10, 16471–16479. [Google Scholar] [CrossRef]
  117. Syed, A.; Battula, H.; Mishra, S.; Jayanty, S. Distinct Tetracyanoquinodimethane Derivatives: Enhanced Fluorescence in Solutions and Unprecedented Cation Recognition in the Solid State. ACS Omega 2021, 6, 3090–3105. [Google Scholar] [CrossRef]
  118. Pakravesh, F.; Izadyar, M.; Arkan, F. Effect of electron donor and acceptor on the photovoltaic properties of organic dyes for efficient dye-sensitized solar cells. Physica B 2021, 609, 412815–412851. [Google Scholar] [CrossRef]
  119. Li, Y.; Liu, J.; Liu, D.; Li, X.; Xu, Y. D-A-π-A based organic dyes for efficient DSSCs: A theoretical study on the role of π-spacer. Comput. Mater. Sci. 2019, 161, 163–176. [Google Scholar] [CrossRef]
  120. Fu, Y.; Li, B.; Liu, H.; Xue, B.; Liu, E. High efficiency dye-sensitized solar cells based on a series of small dye molecules with N-methylcarbazole derivatives as donors. Mater. Chem. Phys. 2020, 239, 121970–121976. [Google Scholar] [CrossRef]
  121. Li, Y.; Li, X.; Xu, Y. Theoretical screening of high-efficiency sensitizers with D-π-A framework for DSSCs by altering promising donor group. Sol. Energy 2020, 196, 146–156. [Google Scholar] [CrossRef]
  122. Kivala, M.; Boudon, C.; Gisselbrecht, J.P.; Seiler, P.; Gross, M.; Diederich, F. A novel reaction of 7,7,8,8-tetracyanoquinodimethane (TCNQ): Charge-transfer chromophores by [2 + 2] cycloaddition with alkynes. Chem. Commun. 2007, 7, 4731–4733. [Google Scholar] [CrossRef]
  123. Correa-Baena, J.-P.; Saliba, M.; Buonassisi, T.; Graetzel, M.; Abate, A.; Tress, W.; Hagfeldt, A. Promises and challenges of perovskite solar cells. Science 2017, 358, 739–744. [Google Scholar] [CrossRef]
  124. Wu, T.; Qin, Z.; Wang, Y.; Wu, Y.; Chen, W.; Zhang, S.; Cai, M.; Dai, S.; Zhang, J.; Liu, J.; et al. The main progress of perovskite solar cells in 2020–2021. Nano-Micro Lett. 2021, 13, 152. [Google Scholar] [CrossRef]
  125. Yang, M.; Zhang, T.; Schulz, P.; Li, Z.; Li, G.; Kim, D.H.; Guo, N.; Berry, J.J.; Zhu, K.; Zhao, Y. Facile fabrication of large-grain CH3NH3PbI3−xBrx films for high-efficiency solar cells via CH3NH3Br-selective Ostwald ripening. Nat. Commun. 2016, 7, 12305–12314. [Google Scholar] [CrossRef]
  126. Koh, T.M.; Wang, H.; Ng, Y.; Bruno, A.; Mhaisalkar, S.; Mathews, N. Halide perovskite solar cells for building integrated photovoltaics: Transforming building facades into power generators. Adv. Mater. 2022, 34, 2104661–2104699. [Google Scholar] [CrossRef] [PubMed]
  127. Jiang, Q.; Tong, J.; Xian, Y.; Kerner, R.; Dunfield, S.; Xiao, H.; Scheidt, R.A.; Kuciauskas, D.; Wang, X.; Hautzinger, M.P.; et al. Surface reaction for efficient and stable inverted perovskite solar cells. Nature 2022, 611, 278–283. [Google Scholar] [CrossRef]
  128. Wagner, P. Silicon solar cells. Microelectron. J. 1988, 19, 37–50. [Google Scholar] [CrossRef]
  129. Sharma, S.; Jain, K.K.; Sharma, A. Solar cells: In Research and Applications. Mater. Sci. Appl. 2015, 6, 1145–1155. [Google Scholar] [CrossRef]
  130. De Wolf, S.; Holovsky, J.; Moon, S.J.; Löper, P.; Niesen, B.; Ledinsky, M.; Haug, F.J.; Yum, J.H.; Ballif, C. Organometallic halide perovskites: Sharp optical absorption edge and its relation to photovoltaic performance. J. Phys. Chem. Lett. 2014, 5, 1035–1039. [Google Scholar] [CrossRef] [PubMed]
  131. Comin, R.; Walters, G.; Thibau, E.S.; Voznyy, O.; Lu, Z.H.; Sargent, E.H. Structural, optical, and electronic studies of wide-bandgap lead halide perovskites. J. Mater. Chem. C. 2015, 3, 8839–8843. [Google Scholar] [CrossRef]
  132. Herz, L.M. Charge-carrier mobilities in Metal Halide perovskites: Fundamental mechanisms and limits. ACS Energy Lett. 2017, 2, 1539–1548. [Google Scholar] [CrossRef]
  133. Wolff, C.M.; Bourelle, S.A.; Phuong, L.Q.; Kurpiers, J.; Feldmann, S.; Caprioglio, P.; Marquez, J.A.; Wolansky, J.; Unold, T.; Stolterfoht, M.; et al. Orders of recombination in complete Perovskite Solar cells—Linking time-resolved and steady-state measurements. Adv. Energy Mater. 2021, 11, 2101823. [Google Scholar] [CrossRef]
  134. Ma, Y.; Gong, J.; Zeng, P.; Liu, M. Recent progress in Interfacial Dipole Engineering for Perovskite Solar cells. Nano-Micro Lett. 2023, 15, 173. [Google Scholar] [CrossRef]
  135. Zhang, C.; Arumugam, G.M.; Liu, C.; Hu, J.; Yang, Y.; Schropp, R.E.J.; Mai, Y. Inorganic halide perovskite materials and solar cells. APL Mater. 2019, 7, 120702. [Google Scholar] [CrossRef]
  136. Wang, Q.; Shao, Y.; Dong, Q.; Xiao, Z.; Yuan, Y.; Huang, J. Large fill-factor bilayer iodine perovskite solar cells fabricated by a low-temperature solution-process. Energy Environ. Sci. 2014, 7, 2359–2365. [Google Scholar] [CrossRef]
  137. Liu, D.; Li, Y.; Yuan, J.; Hong, Q.; Shi, G.; Yuan, D.; Wei, J.; Huang, C.; Tang, J.; Fung, M.-K. Improved performance of inverted planar perovskite solar cells with F4TCNQ doped PEDOT:PSS hole transport layers. J. Mater. Chem. A 2017, 5, 5701–5708. [Google Scholar] [CrossRef]
  138. Huang, L.; Hu, Z.; Xu, J.; Zhang, K.; Zhang, J.; Zhang, J.; Zhu, Y. Efficient and Stable Planar Perovskite Solar Cells with a Non-Hygroscopic Small Molecule Oxidant Doped Hole. Transport Layer. Electrochim. Acta 2016, 196, 328–336. [Google Scholar] [CrossRef]
  139. Song, D.; Wei, D.; Cui, P.; Li, M.; Duan, Z.; Wang, T.; Ji, J.; Li, Y.; Michel Mbengue, J.; Li, Y.; et al. Dual Function Interfacial Layer for Highly Efficient and Stable Lead Halide Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 6091–6097. [Google Scholar] [CrossRef]
  140. Momblona, C.; Gil-Escrig, L.; Bandiello, E.; Hutter, E.M.; Sessolo, M.; Lederer, K.; Blochwitz-Nimoth, J.; Bolink, H.J. Efficient Vacuum Deposited p–i–n and n–i–p Perovskite Solar Cells Employing Doped Charge Transport Layers. Energy Environ. Sci. 2016, 9, 3456–3463. [Google Scholar] [CrossRef]
  141. Zhao, S.; Zhuang, J.; Liu, X.; Zhang, H.; Zheng, R.; Peng, X.; Gong, X.; Guo, H.; Wang, H.; Li, H. F4TCNQ doped strategy of nickel oxide as high-efficient hole transporting materials for invert perovskite solar cell. Mater. Sci. Semicond. Proc. 2001, 121, 105458. [Google Scholar] [CrossRef]
  142. Chen, H.; Wei, Q.; Saidaminov, M.I.; Wang, F.; Johnston, A.; Hou, Y.; Peng, Z.; Xu, K.; Zhou, W.; Liu, Z. Efficient and stable inverted perovskite solar cells incorporating secondary amines. Adv. Mater. 2019, 31, 1903559. [Google Scholar] [CrossRef]
  143. Xu, L.; Chen, X.; Jin, J.; Liu, W.; Dong, B.; Bai, X.; Song, H.; Reiss, P. Inverted perovskite solar cells employing doped NiO hole transport layers: A review. Nano Energy 2019, 63, 103860–103874. [Google Scholar] [CrossRef]
  144. Hu, L.; Zhang, L.; Ren, W.; Zhang, C.; Wu, Y.; Liu, Y.; Sun, Q.; Dai, Z.; Cui, Y.; Cai, L.; et al. High Efficiency Perovskite Solar Cells with PTAA Hole Transport Layer enabled by PMMA:F4TCNQ Buried Interface Layer. J. Mater. Chem. C 2022, 10, 9714–9722. [Google Scholar] [CrossRef]
  145. Jung, M.-C.; Raga, S.R.; Ono, L.K.; Qi, Y. Substantial improvement of perovskite solar cells stability by pinhole-free hole transport layer with doping engineering. Sci. Rep. 2015, 5, 9863. [Google Scholar] [CrossRef]
  146. Calado, P.; Telford, A.M.; Bryant, D.; Li, X.; Nelson, J.; O’Regan, B.C.; Barnes, P.R.F. Evidence for ion migration in hybrid perovskite solar cells with minimal hysteresis. Nat. Commun. 2016, 7, 13831–13841. [Google Scholar] [CrossRef]
  147. Luo, J.; Jia, C.; Wan, Z.; Han, F.; Zhao, B.; Wang, R. The novel dopant for hole-transporting material opens a new processing route to efficiently reduce hysteresis and improve stability of planar perovskite solar cells. J. Power Sources 2017, 342, 886–895. [Google Scholar] [CrossRef]
  148. Trifiletti, V.; Degousée, T.; Manfredi, N.; Fenwick, O.; Colella, S.; Rizzo, A. Molecular Doping for Hole Transporting Materials in Hybrid Perovskite Solar Cells. Metals 2020, 10, 14. [Google Scholar] [CrossRef]
  149. Mahmud, M.A.; Elumalai, N.K.; Upama, M.B.; Wang, D.; Gonçales, V.R.; Wright, M.; Xu, C.; Haque, F.; Uddin, A. A high performance and low-cost hole transporting layer for efficient and stable perovskite solar cells. Phys. Chem. Chem. Phys. 2017, 19, 21033–21045. [Google Scholar] [CrossRef]
  150. Zhang, F.; Song, J.; Hu, R.; Xiang, Y.; He, J.; Hao, Y.; Lian, J.; Zhang, B.; Zeng, P.; Qu, J. Interfacial Passivation of the p-Doped Hole-Transporting Layer Using General Insulating Polymers for High-Performance Inverted Perovskite Solar Cells. Small 2018, 14, 1704007–1704017. [Google Scholar] [CrossRef] [PubMed]
  151. Li, J.; Rochester, C.W.; Jacobs, I.E.; Friedrich, S.; Stroeve, P.; Riede, M.; Moule, A.J. Measurement of Small Molecular Dopant F4TCNQ and C60F36 Diffusion in Organic Bilayer Architectures. ACS Appl. Mater. Interfaces 2015, 7, 28420–28428. [Google Scholar] [CrossRef] [PubMed]
  152. Zhang, H.; Liu, M.; Yang, W.; Judin, L.; Hukka, T.I.; Priimagi, A.; Deng, Z.; Vivo, P. Thionation Enhances the Performance of Polymeric Dopant-Free Hole-Transporting Materials for Perovskite Solar Cells. Adv. Mater. Interfaces 2019, 18, 1901036. [Google Scholar] [CrossRef]
  153. Tran, H.C.V.; Jiang, W.; Lyu, M.; Chae, H. Tetrahydrofuran as solvent for P3HT/F4TCNQ hole-transporting layer to increase the efficiency and stability of FAPbi3-based perovskite solar cell. J. Phys. Chem. C 2020, 124, 14099–14104. [Google Scholar] [CrossRef]
  154. Yao, H.; Hou, J. Recent Advances in Single-Junction Organic Solar Cells. Ang. Chem. Int. Ed. 2022, 61, e202209021–e202209030. [Google Scholar] [CrossRef]
  155. Esqueda, M.L.; Vergara, M.E.S.; Bada, J.R.Á.; Salcedo, R. CuPc: Effects of its doping and a study of its organic-semiconducting properties for application in flexible devices. Materials 2019, 12, 434. [Google Scholar] [CrossRef]
  156. Capitan, M.J.; Alvarez, J.; Naviod, C. Study of the electronic structure of electron accepting cyano-films: TCNQ versus TCNE. Phys. Chem. Chem. Phys. 2018, 20, 10450–10459. [Google Scholar] [CrossRef] [PubMed]
  157. Suzuki, A.; Hasegawa, R.; Funayama, K.; Oku, T.; Okita, M.; Fukunishi, S.; Tachikawa, T.; Hasegawa, T. Additive effects of CuPcX4-TCNQ on CH3NH3PbI3 perovskite solar cells. J. Mater. Sci. Mater. Electron. 2023, 34, 588. [Google Scholar] [CrossRef]
  158. Boix, P.P.; Nonomura, K.; Mathews, N.; Mhaisalkar, S.G. Current Progress and Future Perspectives for Organic/Inorganic Perovskite Solar Cells. Mater. Today 2014, 17, 16–23. [Google Scholar] [CrossRef]
  159. Green, M.A.; Ho-Baillie, A.; Snaith, H.J. The Emergence of Perovskite Solar Cells. Nat. Photonics 2014, 8, 506–514. [Google Scholar] [CrossRef]
  160. Berry, J.; Buonassisi, T.; Egger, D.A.; Hodes, G.; Kronik, L.; Loo, Y.-L.; Lubomirsky, I.; Marder, S.R.; Mastai, Y.; Miller, J.S.; et al. Hybrid Organic–Inorganic Perovskites (HOIPs): Opportunities and Challenges. Adv. Mater. 2015, 27, 5102–5112. [Google Scholar] [CrossRef] [PubMed]
  161. Tiep, N.H.; Ku, Z.; Fan, H.J. Recent Advances in Improving the Stability of Perovskite Solar Cells. Adv. Energy Mater. 2016, 6, 1501420. [Google Scholar] [CrossRef]
  162. Habisreutinger, S.N.; McMeekin, D.P.; Snaith, H.J.; Nicholas, R.J. Research Update: Strategies for Improving the Stability of Perovskite Solar Cells. APL Mater. 2016, 4, 091503–091518. [Google Scholar] [CrossRef]
  163. Hörantner, M.T.; Leijtens, T.; Ziffer, M.E.; Eperon, G.E.; Christoforo, M.G.; McGehee, M.D.; Snaith, H.J. The Potential of Multijunction Perovskite Solar Cells. ACS Energy Lett. 2017, 2, 2506–2513. [Google Scholar] [CrossRef]
  164. Zhang, L.; Zhu, X.; Deng, D.; Wang, Z.; Zhang, Z.; Li, Y.; Zhang, J.; Lv, K.; Liu, L.; Zhang, X.; et al. High Miscibility Compatible with Ordered Molecular Packing Enables an Excellent Efficiency of 16.2% in All-Small-Molecule Organic Solar Cells. Adv. Mater. 2022, 34, 2106316. [Google Scholar] [CrossRef]
  165. Ge, J.; Hong, L.; Ma, H.; Ye, Q.; Chen, Y.; Xie, L.; Song, W.; Li, D.; Chen, Z.; Yu, K.; et al. Asymmetric Substitution of End-Groups Triggers 16.34% Efficiency for All-Small- Molecule Organic Solar Cells. Adv. Mater. 2022, 34, 2202752. [Google Scholar] [CrossRef] [PubMed]
  166. Yan, H.; Manion, J.G.; Yuan, M.; Pelayo García de Arquer, F.; McKeown, G.R.; Beaupré, S.; Leclerc, M.; Sargent, E.H.; Seferos, D.S. Increasing Polymer Solar Cell Fill Factor by Trap-filling with F4TCNQ at Parts Per Thousand Concentration. Adv. Mater. 2016, 28, 6491–6496. [Google Scholar] [CrossRef]
  167. Xiong, Y.; Ye, L.; Gadisa, A.; Zhang, Q.; Rech, J.J.; You, W.; Ade, H. Revealing the Impact of F4-TCNQ as Additive on Morphology and Performance of High-Efficiency Nonfullerene Organic Solar Cells. Adv. Funct. Mater. 2019, 29, 1806262–1806271. [Google Scholar] [CrossRef]
  168. Yang, S.; Zhao, H.; Wu, M.; Yuan, S.; Han, Y.; Liu, Z.; Guo, K.; Liu, S.; Yang, S.; Zhao, H.; et al. Highly efficient and stable planar CsPbI2Br perovskite solar cell with a new sensitive-dopant-free hole transport layer obtained via an effective Surface passivation. Sol. Energy Mater. Sol. Cells 2019, 201, 110052. [Google Scholar] [CrossRef]
  169. Liu, Q.; Cai, W.; Wang, W.; Wang, H.; Zhong, Y.; Zhao, K. Controlling phase transition toward future low-cost and eco-friendly printing of perovskite solar cells. J. Phys. Chem. Lett. 2022, 13, 6503–6513. [Google Scholar] [CrossRef] [PubMed]
  170. Ma, M.; Zhang, C.; Ma, Y.; Li, W.; Wang, Y.; Wu, S.; Liu, C.; Mai, Y. Efficient and Stable Perovskite Solar Cells and Modules Enabled by Tailoring Additive Distribution According to the Film Growth Dynamics. Nano-Micro Lett. 2025, 17, 39–52. [Google Scholar] [CrossRef] [PubMed]
  171. Khan, S.Y.; Rauf, S.; Liu, S.; Chen, W.; Shen, Y.; Kumar, M. Revolutionizing the solar photovoltaic efficiency: A comprehensive review on the cutting-edge thermal management methods for advanced and conventional solar photovoltaics. Energy Environ. Sci. 2025, 18, 1130–1175. [Google Scholar] [CrossRef]
  172. Yan, J.; Savenije, T.J.; Mazzarella, L.; Isabella, O. Progress and challenges on scaling up of perovskite solar cell technology. Sustain. Energy Fuels 2022, 6, 243–266. [Google Scholar] [CrossRef]
  173. Shen, W.; Zhao, Y.; Liu, F. Highlights of mainstream solar cell efficiencies in 2023. Front. Energy 2024, 18, 8–15. [Google Scholar] [CrossRef]
  174. Peng, C.; Ning, G.-H.; Su, J.; Zhong, G.; Tang, W.; Tian, B.; Su, C.; Yu, D.; Zu, L.; Yang, J.; et al. Reversible multi-electron re- dox chemistry of π-conjugated N-containing heteroaromatic molecule-based organic cathodes. Nat. Energy 2017, 2, 17074. [Google Scholar] [CrossRef]
  175. Yang, L.; Chen, J.; Park, S.; Wang, H. Recent progress on metal-organic framework derived carbon and its composites as anode materials for potassium-ion batteries. Energy Mater. 2023, 3, 300042–300074. [Google Scholar] [CrossRef]
  176. Lee, J.; Lee, H.; Bak, C.; Hong, Y.; Joung, D.; Ko, J.B.; Lee, Y.M.; Kim, C. Enhancing hydrophilicity of thick electrodes for high energy density aqueous batteries. Nano-Micro Lett. 2023, 15, 97–108. [Google Scholar] [CrossRef]
  177. Lv, C.; Xu, W.; Liu, H.; Zhang, L.; Chen, S.; Yang, X.; Xu, X.; Yang, D. 3D Sulfur and Nitrogen Co doped Carbon Nanofiber Aerogels with Optimized Electronic Structure and Enlarged Interlayer Spacing Boost Potassium-Ion Storage. Small 2019, 15, 1900816–1900824. [Google Scholar] [CrossRef]
  178. Li, D.; Guo, Y.; Zhang, C.; Chen, X.; Zhang, W.; Mei, S.; Yao, C.-J. Unveiling Organic Electrode Materials in Aqueous Zinc-Ion Batteries: From Structural Design to Electrochemical Performance. Nano-Micro Lett. 2024, 16, 194–228. [Google Scholar] [CrossRef]
  179. Hanyu, Y.; Honma, I. Rechargeable quasi-solid state lithium battery with organic crystalline cathode. Sci. Rep. 2012, 2, 453–458. [Google Scholar] [CrossRef]
  180. Lee, S.; Hong, J.; Jung, S.K.; Ku, K.; Kwon, G.; Seong, W.M.; Kim, H.; Yoon, G.; Kang, I.; Hong, K.; et al. Charge-transfer complexes for high-power organic rechargeable batteries. Energy Storage Mater. 2019, 20, 462–469. [Google Scholar] [CrossRef]
  181. Kye, H.; Kang, Y.; Jang, D.; Kwon, J.E.; Kim, B.-G. p-Type Redox-Active Organic Electrode Materials for Next-Generation Rechargeable Batteries. Adv. Energy Sustain. Res. 2022, 3, 2200030–2200055. [Google Scholar] [CrossRef]
  182. Shi, H.J.; Ye, Y.J.; Liu, K.; Song, Y.; Sun, X. A Long-Cycle-Life Self-Doped Polyaniline Cathode for Rechargeable Aqueous Zinc Batteries. Angew. Chem. Int. Ed. 2018, 57, 16359–16363. [Google Scholar] [CrossRef] [PubMed]
  183. Yi, J.; Liang, P.; Liu, X.; Wu, K.; Liu, Y.; Wang, Y. Challenges, mitigation strategies and perspectives in development of zinc-electrode materials and fabrication for rechargeable zinc–air batteries. Energy Environ. Sci. 2018, 11, 3075–3095. [Google Scholar] [CrossRef]
  184. Li, H.; Fang, M.; Hou, Y.; Tang, R.; Yang, Y.; Zhong, C.; Li, Q.; Li, Z. Different Effect of the Additional Electron-Withdrawing Cyano Group in Different Conjugation Bridge: The Adjusted Molecular Energy Levels and Largely Improved Photovoltaic Performance. ACS Appl. Mater. Interface 2016, 8, 12134–12140. [Google Scholar] [CrossRef]
  185. Casey, A.; Dimitrov, S.D.; Shakya-Tuladhar, P.; Fei, Z.; Nguyen, M.; Han, Y.; Anthopoulos, T.D.; Durrant, J.R.; Heeney, M. Effect of Systematically Tuning Conjugated Donor Polymer Lowest Unoccupied Molecular Orbital Levels via Cyano Substitution on organic Photovoltaic Device Performance. Chem. Mater. 2016, 28, 5110–5120. [Google Scholar] [CrossRef]
  186. Ji, Z.; Dong, H.; Liu, M.; Hu, W. Water-controlled synthesis of low-dimentuinal Molecular crystals and the fabrication of a new water and moisture indicator. Nano Res. 2009, 11, 857–864. [Google Scholar] [CrossRef]
  187. Kundu, D.; Oberholzer, P.; Glaros, C.; Bouzid, A.; Tervoort, E.; Pasquarello, A.; Niederberger, M. Organic cathode for aqueous Zn-ion batteries: Taming a unique phase evolution toward stable electrochemical cycling. Chem. Mater. 2018, 30, 3874–3881. [Google Scholar] [CrossRef]
  188. Chola, N.M.; Nagarale, R.K. TCNQ confined in porous organic structure as cathode for aqueous zinc battery. J. Electrochem. Soc. 2020, 167, 100552–100561. [Google Scholar] [CrossRef]
  189. Wang, S.; Deng, W.; Geng, Z.; Li, P.; Hu, N.; Zhu, L.; Sun, W.; Li, C.M. Significantly raising tetracyanoquinodimethane electrode performance in zinc-ion battery at low temperatures by eliminating impurities. Battery Energy 2023, 2, 20220050–20220073. [Google Scholar] [CrossRef]
  190. Wang, S.; Sun, Z.; Zhou, Y.; Deng, W. Effect of Fluoride Atoms on the Electrochemical Performance of Tetracyanoquinodimethane (TCNQ) Electrodes in Zinc-ion Batteries. Chem. Phys. Chem. 2023, 24, e202300436–e202300441. [Google Scholar] [CrossRef]
  191. Canevet, D.; Sallé, M.; Zhang, G.; Zhang, D.; Zhu, D. Tetrathiafulvalene (TTF) derivatives: Key building-blocks for switchable processes. Chem. Commun. 2009, 17, 2245–2269. [Google Scholar] [CrossRef]
  192. Ma, W.; Lei, C.; Liu, T.; Liang, X. A high-voltage tetracyanoquinodimethane (TCNQ) organic cathode enabled by Na+ and I synergy for durable aqueous batteries. Sci. China Chem. 2025, 68, 4478–4485. [Google Scholar] [CrossRef]
  193. Lee, S.; Kim, J.; Hong, J.; Kang, K. Unlocking the Full Redox Capability of Organic Charge-Transfer Complex in High-Loading Electrodes for Organic Rechargeable Batteries. Adv. Energy Mater. 2025, 15, 2404116–2404124. [Google Scholar] [CrossRef]
  194. Wang, Z.L.; Xu, D.; Xu, J.J.; Zhang, X.B. Oxygen electrocatalysts in metal-air batteries: From aqueous to nonaqueous electrolytes. Chem. Soc. Rev. 2014, 43, 7746–7786. [Google Scholar] [CrossRef]
  195. Balogun, M.-S.; Wu, Z.; Luo, Y.; Qiu, W.; Fan, X.; Long, B.; Huang, M.; Liu, P.; Tong, Y. High power density nitridated hematite (α-Fe2O3) nanorods as anode for high-performance flexible lithium ion batteries. J. Power Sources 2016, 308, 7–17. [Google Scholar] [CrossRef]
  196. Huang, A.; Ma, Y.; Peng, J.; Li, L.; Chou, S.-L.; Ramakrishna, S.; Peng, S. Tailoring the structure of silicon-based materials for lithium-ion batteries via electrospinning technology. eScience 2021, 1, 141–162. [Google Scholar] [CrossRef]
  197. Luo, D.; Li, M.; Zheng, Y.; Ma, Q.; Gao, R.; Zhang, Z.; Dou, H.; Wen, G.; Shui, L.; Yu, A.; et al. Electrolyte design for lithium metal anode-based batteries toward extreme temperature application. Adv. Sci. 2021, 8, 2101051–2101070. [Google Scholar] [CrossRef]
  198. Gou, X.; Hao, Z.; Hao, Z.; Yang, G.; Yang, Z.; Zhang, X.; Yan, Z.; Zhao, Q.; Chen, J. In situ surface self-reconstruction strategies in Li-rich Mn-based layered cathodes for energy-dense Li-ion batteries. Adv. Funct. Mater. 2022, 32, 2112088–2112103. [Google Scholar] [CrossRef]
  199. Chen, Z.; Mo, F.; Wang, T.; Yang, Q.; Huang, Z.; Wang, D.; Liang, G.; Chen, A.; Li, Q.; Guo, Y.; et al. Zinc/selenium conversion battery: A system highly compatible with both organic and aqueous electrolytes. Energy Environ. Sci. 2021, 14, 2441–2450. [Google Scholar] [CrossRef]
  200. Zhu, C.; Zhou, J.; Wang, Z.; Zhou, Y.; He, X.; Zhou, X.; Liu, J.; Yan, C.; Qian, T. Phase diagrams guided design of low-temperature aqueous electrolyte for Zn metal batteries. Chem. Eng. J. 2023, 454, 140413–140419. [Google Scholar] [CrossRef]
  201. Li, Z.; Liao, Y.; Wang, Y.; Cong, J.; Ji, H.; Huang, Z.; Huang, Y. A co-solvent in aqueous electrolyte towards ultra long-life rechargeable zinc ion batteries. Energy Storage Mater. 2023, 56, 174–182. [Google Scholar] [CrossRef]
  202. Fang, C.; Huang, Y.; Yuan, L.; Liu, Y.; Chen, W.; Huang, Y.; Chen, K.; Han, J.; Liu, Q.; Huang, Y. A metal-organic compound as cathode material with super high capacity achieved by reversible cationic and anionic redox chemistry for high-energy sodium-ion batteries. Ang. Chem. Int. Ed. 2017, 56, 6793–6797. [Google Scholar] [CrossRef] [PubMed]
  203. Precht, R.; Hausbrand, R.; Jaegermann, W. Electronic structure and electrode properties of tetracyanoquinodimethane (TCNQ): A surface science investigation of lithium intercalation into TCNQ. Phys. Chem. Chem. Phys. 2015, 17, 6588–6596. [Google Scholar] [CrossRef]
  204. Ma, E.; Zhou, C.; Fan, B.; Wu, C.; Li, Z.; Lu, H.; Li, J. Endowing CuTCNQ with a new role: A high-capacity cathode for K-ion batteries. Chem. Commun. 2018, 54, 5578–5581. [Google Scholar] [CrossRef]
  205. Huang, Y.; Fang, C.; Zhang, W.; Liu, Q.J.; Huang, Y.H. Room temperature N-heterocyclic carbene manganese catalyzed selective N-alkylation of anilines with alcohols. Chem. Commun. 2019, 55, 608–611. [Google Scholar] [CrossRef]
  206. Fujihara, Y.; Kobayashi, H.; Takaishi, S.; Tomai, S.; Yamashita, M.; Honma, I. Electrical Conductivity-Relay between Organic Charge-Transfer and radical Salts toward Conductive Additive-Free rechargeable Battery. ACS Appl. Mater. Interfaces 2020, 12, 25748–25755. [Google Scholar] [CrossRef] [PubMed]
  207. Whittingham, M.S. Lithium Batteries and cathode materials. Chem. Rev. 2004, 104, 4271–4301. [Google Scholar] [CrossRef]
  208. Kobayashi, H.; Hibino, M.; Ogasawara, Y.; Yamaguchi, K.; Kudo, T.; Okuoka, S.-I.; Yonehara, K.; Ono, H.; Sumida, Y.; Oshima, M. Improved performance of Co-doped Li2O cathodes for lithium-peroxide batteries using LiCoO2 as a dopant source. J. Power Sources 2016, 306, 567–572. [Google Scholar] [CrossRef]
  209. Pender, J.P.; Jha, G.; Youn, D.H.; Ziegler, J.M.; Andoni, I.; Choi, E.J.; Heller, A.; Dunn, B.S.; Weiss, P.S.; Penner, R.M.; et al. Electrode Degradation in Lithium-Ion Batteries. ACS Nano 2020, 14, 1243–1295. [Google Scholar] [CrossRef]
  210. Heintz, R.A.; Zhao, H.H.; Ouyang, X.; Grandinetti, G.; Cowen, J.; Dunbar, K.R. New Insight into the Nature of Cu(TCNQ): Solution Routes to Two Distinct Polymorphs and Their Relationship to Crystalline Films That Display Bistable Switching Behavior. Inorg. Chem. 1999, 38, 144–156. [Google Scholar] [CrossRef]
  211. Song, H.W.; Gong, Y.; Su, J.; Li, W.Y.; Li, Y.; Gu, L.; Wang, C.H. Surfaces/Interfaces Modification for Vacancies Enhancing Lithium Storage Capability of Cu2O Ultrasmall Nanocrystals. ACS Appl. Mater. Interfaces 2018, 10, 35137–35144. [Google Scholar] [CrossRef]
  212. Leong, C.F.; Usov, P.M.; D′Alessandro, D.M. Intrinsically conducting metal–organic frameworks. MRS Bull. 2016, 41, 858–864. [Google Scholar] [CrossRef]
  213. Yin, D.; Wang, Z.; Li, Q.; Xue, H.; Cheng, Y.; Wang, L.; Huang, G. In Situ Growth of Lithiophilic MOF Layer Enabling Dendrite-free Lithium Deposition. iScience 2020, 23, 101869–101879. [Google Scholar] [CrossRef]
  214. Schoedel, A.; Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O.M. Structures of Metal-Organic Frameworks with Rod Secondary Building Units. Chem. Rev. 2016, 116, 12466–12535. [Google Scholar] [CrossRef]
  215. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef]
  216. Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H. Commercial metal–organic frameworks as heterogeneous catalysts. Chem. Commun. 2012, 48, 11275–11288. [Google Scholar] [CrossRef] [PubMed]
  217. Placke, T.; Heckmann, A.; Schmuch, R.; Meister, P.; Beltrop, K.; Winter, M. Perspective on Performance, Cost, and Technical Challenges for Practical Dual-Ion Batteries. Joule 2018, 2, 2528–2550. [Google Scholar] [CrossRef]
  218. Rodriguez-Perez, I.A.; Ji, X.L. Anion Hosting Cathodes in Dual-Ion Batteries. ACS Energy Lett. 2017, 2, 1762–1770. [Google Scholar] [CrossRef]
  219. Wang, M.; Tang, Y.B. A review on the features and progress of dual-ion batteries. Adv. Energy Mater. 2018, 8, 1703320–1703339. [Google Scholar] [CrossRef]
  220. Dühnen, S.; Nölle, R.; Wrogemann, J.; Winter, M.; Placke, T. Reversible Anion Storage in a Metal-Organic Framework for Dual-Ion Battery Systems. J. Electrochem. Soc. 2019, 166, A5474–A5482. [Google Scholar] [CrossRef]
  221. Wu, F.X.; Yushin, G. Conversion cathodes for rechargeable lithium and lithium-ion batteries. Energy Environ. Sci. 2017, 10, 435–459. [Google Scholar] [CrossRef]
  222. Jia, H.P.; Kloepsch, R.; He, X.; Evertz, M.; Nowak, S.; Li, J.; Winter, M.; Placke, T. Nanostructured ZnFe2O4 as Anode Material for Lithium-Ion Batteries: Ionic Liquid-Assisted Synthesis and Performance Evaluation with Special Emphasis on Comparative Metal Dissolution. Acta Chim. Slov. 2016, 63, 470–483. [Google Scholar] [CrossRef]
  223. Qian, Y.X.; Niehoff, P.; Zhou, D.; Adam, R.; Mikhailova, D.; Pyschik, M.; Borner, M.; Kloepsch, R.; Rafaja, D.; Schumacher, G.; et al. Investigation of Nano-Sized Cu(II)O As a High Capacity Conversion Material for Li Metal Cells and Lithium Ion Full Cells. J. Mater. Chem. A 2017, 5, 6556–6568. [Google Scholar] [CrossRef]
  224. Wei, W.; Li, L.; Zhang, L.; Hong, J.; He, G. An all-solid-state Li organic battery with Quinone-based polymer cathode and composite polymer electrolyte. Electrochem. Commun. 2018, 90, 21–25. [Google Scholar] [CrossRef]
  225. Huang, W.; Zheng, S.; Zhang, X.; Zhou, W.; Xiong, W.; Chen, J. Synthesis and application of Calix[6]quinone as a High-Capacity Organic Cathode for Plastic Crystal Electrolyte-Based Lithium-Ion Batteries. Energy Storage Mater. 2020, 26, 465–471. [Google Scholar] [CrossRef]
  226. Hao, F.; Liang, Y.; Zhang, Y.; Chen, Z.; Zhang, J.; Ai, Q.; Guo, H.; Fan, Z.; Lou, J.; Yao, Y. High-energy all-solid-state organic–lithium batteries based on ceramic electrolytes. ACS Energy Lett. 2021, 6, 201–207. [Google Scholar] [CrossRef]
  227. Bonilla, M.R.; Daza, F.A.G.; Ranque, P.; Aguesse, F.; Carrasco, J.; Akhmatskaya, E. Unveiling interfacial Li-ion dynamics in Li7La3Zr2O12/PEO(LiTFSI) composite polymer-ceramic solid electrolytes for all-solid-state lithium batteries. ACS Appl. Mater. Interfaces 2021, 13, 30653–30667. [Google Scholar] [CrossRef] [PubMed]
  228. An, Y.; Han, X.; Liu, Y.; Azhar, A.; Na, J.; Nanjundan, A.K.; Wang, S.; Yu, J.; Yamauchi, Y. Progress in Solid Polymer Electrolytes for Lithium-Ion Batteries and Beyond. Small 2022, 18, 2103617–2103652. [Google Scholar] [CrossRef]
  229. Canals-Riclot, J.; Irié-Bi, B.; Dolhem, F.; Becuwe, M.; Gautron, E.; Seznec, V.; Dedryvère, R. Organic All-Solid-State Lithium Metal Battery Using Polymer/Covalent Organic Framework Electrolyte. Batter. Supercaps 2024, 8, e202400357–e202400364. [Google Scholar] [CrossRef]
  230. Xia, J.; Wang, R.; Qian, C.; Sun, K.; Liu, H.; Guo, C.; Li, J.; Yu, F.; Bao, W. Supercapacitors of Nanocrystalline Covalent Organic Frameworks—A Review. Crystals 2022, 12, 1350. [Google Scholar] [CrossRef]
  231. Sasmal, H.S.; Mahato, A.K.; Majumder, P.; Banerjee, R. Landscaping Covalent Organic Framework Nanomorphologies. Am. Chem. Soc. 2022, 144, 11482–11498. [Google Scholar] [CrossRef]
  232. Wu, Y.; Yan, D.; Zhang, Z.; Matsushita, M.M.; Awaga, K. Electron Highways into Nanochannels of Covalent Organic Frameworks for High Electrical Conductivity and Energy Storage. ACS Appl. Mater. Interfaces 2019, 11, 7661–7665. [Google Scholar] [CrossRef]
  233. Wang, X.; Wang, Y.; Qi, H.; Chen, Y.; Guo, W.; Yu, H.; Chen, H.; Ying, Y. Humidity-Independent Artificial Olfactory Array Enabled by Hydrophobic Core-Shell Dye/MOFs@COFs. Composites for Plant Disease Diagnosis. ACS Nano 2022, 16, 14297–14307. [Google Scholar] [CrossRef] [PubMed]
  234. Li, Y.; Zhao, L.; Bai, Y.; Feng, F. Applications of covalent organic frameworks (COFs)-based sensors for food safety: Synthetic strategies, characteristics and current state-of-art. Food Chem. 2025, 469, 142495. [Google Scholar] [CrossRef] [PubMed]
  235. Yang, Y.; Mallick, S.; Izquierdo-Ruiz, F.; Schäfer, C.; Xing, X.; Rahm, M.; Börjesson, K. A Highly Conductive All-Carbon Linked 3D Covalent Organic Framework Film. Small 2021, 17, 2103152–2103157. [Google Scholar] [CrossRef] [PubMed]
  236. Slater, M.D.; Kim, D.; Lee, E.; Johnson, C.S. Sodium-Ion Batteries. Adv. Funct. Mater. 2013, 23, 947–958. [Google Scholar] [CrossRef]
  237. Pan, H.; Hu, Y.S.; Chen, L. Room-temperature stationary sodium-ion batteries for large-scale electric energy storage. Energy Environ. Sci. 2013, 6, 2338–2360. [Google Scholar] [CrossRef]
  238. Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research development on sodium-ion batteries. Chem. Rev. 2014, 114, 11636–11682. [Google Scholar] [CrossRef]
  239. Dhir, S.; Wheeler, S.; Capone, I.; Pasta, M. Outlook on K-Ion Batteries. Chem 2020, 6, 2442–2460. [Google Scholar] [CrossRef]
  240. Roberts, S.; Kendrick, E. The reemergence of sodium ion batteries: Testing, processing, and manufacturability. Nanotechnol. Sci. Appl. 2018, 11, 23–33. [Google Scholar] [CrossRef]
  241. Zhang, J.; Liu, T.; Cheng, X.; Xia, M.; Zheng, R.; Peng, N.; Yu, H.; Shui, M.; Shu, J. Development status and future prospect of non-aqueous potassium ion batteries for large scale energy storage. Nano Energy 2019, 60, 340–361. [Google Scholar] [CrossRef]
  242. Hosaka, T.; Kubota, K.; Hameed, A.S.; Komaba, S. Research development on K-ion batteries. Chem. Rev. 2020, 120, 6358–6466. [Google Scholar] [CrossRef]
  243. Zhang, S.; Liu, Y.; Fan, Q.; Zhang, C.; Zhou, T.; Kalantar-Zadeh, K.; Guo, Z. Liquid metal batteries for future energy storage. Energy Environ. Sci. 2021, 14, 4177–4202. [Google Scholar] [CrossRef]
  244. Zhang, S.; Fan, Q.; Liu, Y.; Xi, S.; Liu, X.; Wu, Z.; Hao, J.; Pang, W.K.; Zhou, T.; Guo, Z. Dehydration-Triggered Ionic Channel Engineering in Potassium Niobate for Li/K-Ion Storage. Adv. Mater. 2020, 32, 2000380–2000389. [Google Scholar] [CrossRef] [PubMed]
  245. Guo, F.; Huang, Z.; Wang, M.; Song, W.-L.; Lv, A.; Han, X.; Tu, J.; Jiao, S. Active cyano groups to coordinate AlCl2+ cation for rechargeable aluminum batteries. Energy Storage Mater. 2020, 33, 250–257. [Google Scholar] [CrossRef]
  246. Lin, M.C.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.Y.; Guan, M.; Angell, M.; Chen, C.; Yang, J.; Hwang, B.J.; et al. An ultrafast rechargeable aluminium-ion battery. Nature 2015, 520, 325–328. [Google Scholar] [CrossRef]
  247. Yang, H.; Li, H.; Li, J.; Sun, Z.; He, K.; Cheng, H.M.; Li, F. The rechargeable aluminum battery: Opportunities and challenges. Ang. Chem. Int. Ed. 2019, 58, 11978–11996. [Google Scholar] [CrossRef]
  248. Sun, H.; Wang, W.; Yu, Z.; Yuan, Y.; Wang, S.; Jiao, S. A new aluminium-ion battery with high voltage, high safety and low cost. Chem. Commun. 2015, 51, 11892–11895. [Google Scholar] [CrossRef]
  249. Kravchyk, K.V.; Wang, S.; Piveteau, L.; Kovalenko, M.V. Efficient aluminum chloride—Natural graphite battery. Chem. Mater. 2017, 29, 4484–4492. [Google Scholar] [CrossRef]
  250. Wang, D.Y.; Wei, C.Y.; Lin, M.C.; Pan, C.J.; Chou, H.L.; Chen, H.A.; Gong, M.; Wu, Y.; Yuan, C.; Angell, M.; et al. Advanced rechargeable aluminium ion battery with a high-quality natural graphite cathode. Nat. Commun. 2017, 8, 14283–14289. [Google Scholar] [CrossRef]
  251. Zhang, E.; Wang, B.; Wang, J.; Ding, H.; Zhang, S.; Duan, H.; Yu, X.; Lu, B. Rapidly synthesizing interconnected carbon nano cage by microwave toward high-performance aluminum batteries. Chem. Eng. J. 2020, 389, 124407–124414. [Google Scholar] [CrossRef]
  252. Moriyama, K.; Kuramochi, M.; Fujii, K.; Morita, T.; Togo, H. Nitroxyl-radical-catalyzed oxidative coupling of amides with silylated nucleophiles through N-halogenation. Ang. Chem. Int. Ed. 2016, 55, 14546–14551. [Google Scholar] [CrossRef]
  253. Precht, R.; Stolz, S.; Mankel, E.; Mayer, T.; Jaegermann, W.; Hausbrand, R. Investigation of sodium insertion into tetracyanoquinodimethane (TCNQ): Results for a TCNQ thin film obtained by a surface science approach. Phys. Chem. Chem. Phys. 2016, 18, 3056–3064. [Google Scholar] [CrossRef] [PubMed]
  254. Lin, Z.; Zhang, B.; Guo, H.; Wub, Z.; Zoub, H.; Yangc, J.; Wang, Z.L. Super-robust and frequency multiplied triboelectric nanogenerator for efficient harvesting water and wind energy. Nano Energy 2019, 64, 103908–103914. [Google Scholar] [CrossRef]
  255. Shi, B.; Liu, Z.; Zheng, Q.; Meng, J.; Ouyang, H.; Zou, Y.; Jiang, D.; Qu, X.; Yu, M.; Zhao, L.; et al. Body-integrated self-powered system for wearable and implantable applications. ACS Nano 2019, 13, 6017–6024. [Google Scholar] [CrossRef]
  256. Zhou, Q.; Pan, J.; Deng, S.; Xia, F.; Kim, T. Triboelectric Nanogenerator-Based Sensor Systems for Chemical or Biological Detection. Adv. Mater. 2021, 33, 2008276–2008296. [Google Scholar] [CrossRef]
  257. Song, Y.; Wang, N.; Hu, C.; Wang, Z.L.; Yang, Y. Soft triboelectric nanogenerators for mechanical energy scavenging and self-powered sensors. Nano Energy 2021, 84, 105919–105936. [Google Scholar] [CrossRef]
  258. Pu, X.; An, S.; Tang, Q.; Guo, H.; Hu, C. Wearable triboelectric sensors for biomedical monitoring and human-machine interface. iScience 2021, 24, 102027–102047. [Google Scholar] [CrossRef]
  259. Ning, C.; Cheng, R.; Jiang, Y.; Sheng, F.; Yi, J.; Shen, S.; Zhang, Y.; Peng, X.; Dong, K.; Wang, Z.L. Helical fiber strain sensors based on triboelectric nanogenerators for self-powered human respiratory monitoring. ACS Nano 2022, 16, 2811–2821. [Google Scholar] [CrossRef] [PubMed]
  260. Pu, X.; Zhang, C.; Wang, Z.L. Triboelectric nanogenerators as wearable power sources and self-powered sensors. Nat. Sci. Rev. 2023, 10, 170–190. [Google Scholar] [CrossRef]
  261. Fu, Q.; Cui, C.; Meng, L.; Hao, S.; Dai, R.; Yang, J. Emerging cellulose-derived materials: A promising platform for the design of flexible wearable sensors toward health and environment monitoring. Mater. Chem. Front. 2021, 5, 2051–2091. [Google Scholar] [CrossRef]
  262. Wang, P.; Hu, M.; Wang, H.; Chen, Z.; Feng, Y.; Wang, J.; Ling, W.; Huang, Y. The evolution of flexible electronics: From nature, beyond nature, and to nature. Adv. Sci. 2020, 7, 2001116–2001144. [Google Scholar] [CrossRef] [PubMed]
  263. Wang, Z.L.; Wang, A.C. On the origin of contact-electrification. Mater Today 2019, 30, 34–51. [Google Scholar] [CrossRef]
  264. Zhang, Z.; Jiang, D.; Zhao, J.; Liu, G.; Bu, T.; Zhang, C.; Wang, Z.L. Tribovoltaic effect on metal–semiconductor interface for direct-current low-impedance triboelectric nanogenerators. Adv. Energy Mater. 2020, 10, 1903713–1903720. [Google Scholar] [CrossRef]
  265. Slabov, V.; Kopyl, S.; Soares dos Santos, M.P.; Kholkin, A.L. Natural and eco-friendly materials for triboelectric energy harvesting. Nano-Micro Lett. 2020, 12, 42–59. [Google Scholar] [CrossRef]
  266. Jiang, W.; Li, H.; Liu, Z.; Li, Z.; Tian, J.; Shi, B.; Zou, Y.; Ouyang, H.; Zhao, C.; Zhao, L.; et al. Fully bioabsorbable natural-materials-based triboelectric nanogenerators. Adv. Mater. 2018, 30, 1801895–1801904. [Google Scholar] [CrossRef]
  267. Jie, Y.; Jia, X.; Zou, J.; Chen, Y.; Wang, N.; Wang, Z.L.; Cao, X. Natural leaf made triboelectric nanogenerator for harvesting environmental mechanical energy. Adv. Energy Mater. 2018, 8, 1703133–1703139. [Google Scholar] [CrossRef]
  268. Chao, S.; Ouyang, H.; Jiang, D.; Fan, Y.; Li, Z. Triboelectric nanogenerator based on degradable materials. EcoMat 2020, 3, e12072–e12090. [Google Scholar] [CrossRef]
  269. Peng, X.; Dong, K.; Zhang, Y.; Wang, L.; Wei, C.; Lv, T.; Wang, Z.L.; Wu, Z. Sweat-Permeable, Biodegradable, Transparent and Self-powered Chitosan-Based Electronic Skin with Ultrathin Elastic Gold Nanofibers. Adv. Funct. Mater. 2022, 32, 2112241–2112250. [Google Scholar] [CrossRef]
  270. Wang, L.; Wang, Y.; Bo, X.; Wang, H.; Yang, S.; Tao, X.; Zi, Y.; Yu, W.W.; Li, W.J.; Daoud, W.A. High-Performance Biomechanical Energy Harvester Enabled by Switching Interfacial Adhesion via Hydrogen Bonding and Phase Separation. Adv. Funct. Mater. 2022, 32, 2204304–2204314. [Google Scholar] [CrossRef]
  271. Zhang, T.; Wang, L.; Ding, W.; Zhu, Y.; Qian, H.; Zhou, J.; Chen, Y.; Li, J.; Li, W.; Huang, L.; et al. Rationally Designing High-Performance Versatile Organic Memristors through Molecule-Mediated Ion Movements. Adv. Mater. 2023, 35, 2302863–2302874. [Google Scholar] [CrossRef]
  272. Talin, A.A.; Stavila, V.; Allendorf, M.D.; Foster, M.E.; He, Y.; Leonard, F.; Spataru, C.D.; Jones, R.E. Molecule@MOF: A New Class of Opto-electronic Materials. Sandia Rep. 2017, 1, 1–27. [Google Scholar] [CrossRef]
  273. Talin, A.A.; Centrone, A.; Ford, A.C.; Foster, M.E.; Stavila, V.; Haney, P.; Kinney, R.A.; Szalai, V.; Gabaly, F.E.; Yoon, H.P.; et al. Tunable Electrical Conductivity in metal-Organic Framework Thin-Film Devices. Science 2014, 343, 66–69. [Google Scholar] [CrossRef]
  274. Choi, S.; Tan, S.H.; Li, Z.; Kim, Y.; Choi, C.; Chen, P.Y.; Yeon, H.; Yu, S.; Kim, J. SiGe epitaxial memory for neuromorphic computing with reproducible high performance based on engineered dislocations. Nat. Mater. 2018, 17, 335. [Google Scholar] [CrossRef] [PubMed]
Figure 1. TCNQ electron acceptors as stable aromatic radical systems.
Figure 1. TCNQ electron acceptors as stable aromatic radical systems.
Sustainability 17 10612 g001
Scheme 1. Physicochemical properties of TCNQ acceptors for sustainable battery construction.
Scheme 1. Physicochemical properties of TCNQ acceptors for sustainable battery construction.
Sustainability 17 10612 sch001
Figure 2. Properties of TCNQ materials for high-performing and sustainable solar cells and batteries.
Figure 2. Properties of TCNQ materials for high-performing and sustainable solar cells and batteries.
Sustainability 17 10612 g002
Figure 3. The most important discoveries in the solar cell industry, including the contribution of TCNQ [72,73,74,75].
Figure 3. The most important discoveries in the solar cell industry, including the contribution of TCNQ [72,73,74,75].
Sustainability 17 10612 g003
Figure 4. Doping scheme of P3HT with F4TCNQ and effects of charge transfer between these donors and acceptors.
Figure 4. Doping scheme of P3HT with F4TCNQ and effects of charge transfer between these donors and acceptors.
Sustainability 17 10612 g004
Scheme 2. Formation of DADQ compounds.
Scheme 2. Formation of DADQ compounds.
Sustainability 17 10612 sch002
Figure 5. General schematic representation of a perovskite solar cell: (a) without a hole transport layer; (b) with a hole transport layer doped with F4TCNQ.
Figure 5. General schematic representation of a perovskite solar cell: (a) without a hole transport layer; (b) with a hole transport layer doped with F4TCNQ.
Sustainability 17 10612 g005
Figure 6. Schematic illustration of the construction of a stable organic TCNQ all-solid-state lithium metal battery.
Figure 6. Schematic illustration of the construction of a stable organic TCNQ all-solid-state lithium metal battery.
Sustainability 17 10612 g006
Scheme 3. Illustration of physicochemical properties of TCNQ cathode materials for the construction of various sustainable batteries.
Scheme 3. Illustration of physicochemical properties of TCNQ cathode materials for the construction of various sustainable batteries.
Sustainability 17 10612 sch003
Figure 7. Future prospects of TCNQ-based TENGs.
Figure 7. Future prospects of TCNQ-based TENGs.
Sustainability 17 10612 g007
Table 1. Photovoltaic characteristics of solar cells doped with TCNQ or its F4TCNQ derivative.
Table 1. Photovoltaic characteristics of solar cells doped with TCNQ or its F4TCNQ derivative.
TCNQ-Based Organic Solar CellVoc (V)Jsc (mA/cm)FF (%)PCE (%)Ref.
ZnPc/TCNQ
CuPc/TCNQ
0.58
0.48
0.27·10−3
0.37·10−3
18.0
16.0
1.6·10−5
2.8·10−5
[100]
Graphene/TCNQ/graphene0.608.9048.02.58[101]
Spiro-MeOTAD/DMC/F4TCNQ0.9824.0062.414.40[145]
P3HT/F4TCNQ0.879.8563.25.83[103]
DADQs/TCNQ/TiO23.009.1259.011.75[60]
Spiro-MeOTAD/F4TCNQ1.0419.4069.914.30[139]
Spiro-MeOTAD/F4TCNQ0.9518.7256.810.59[154]
CH3NH3PbI3/F4TCNQ1.0620.3075.418.10[140]
PBDTTT-EFT/F4TCNQ0.8017.3961.88.60[164]
PEDOT:PSS/F4TCNQ1.0221.9377.017.22[137]
PTAA/F4TCNQ/Bphen/Al1.1223.3877.120.16[150]
FTAZ:IT-M/F4TCNQ0.9518.7071.012.40[165]
2mF-X59/spiro-OMeTAD/TCNQ/CsPbI2Br1.2014.7875.8513.42[166]
F4TCNQ HTL/NiOx1.0220.0774.515.70[141]
AQ/TCNQ/TiO2
TQ/TCNQ/TiO2
3.25
2.72
10.63
11.86
95.0
94.0
18.92
18.90
[118]
PMMA/F4TCNQ1.0621.8176.317.90[144]
CuPc(NH2)4/TCNQ/CH3NH3PbI30.7723.161.8410.90[157]
Abbreviations: bphen—7-diphenyl-1,10-phenanthroline; DMC—decamethylcobaltocene; FTAZ:IT-M—a high-performance polymer/small-molecule blend; 2mF-X59—N′,N′,N″,N″-tetrakis(4-methoxyphenyl)spiro[fluorene-9,9′-xanthene]-2,7-diamine; PBDTTT-EFT—poly[4,8-bis(5-(2-ethylhexyl)thiophen-2-yl)benzo[1,2-b;4,5-b′]dithiophene-2,6-diyl-alt-(4-(2-ethylhexyl)-3-fluorothieno[3,4-3b]thiophene-)-2-carboxylate-2-6-diyl)]; PTAA—poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine.
Table 2. Features of sustainable TCNQ-based cathode materials for rechargeable batteries.
Table 2. Features of sustainable TCNQ-based cathode materials for rechargeable batteries.
Cathode MaterialType of BatteryAverage Voltage (V)Cycling Stability (%)/CyclesSpecific Capacity (mAh g−1)Ref.
CuTCNQLi+0.01–3.2075/500280.9[50]
CuTCNQLi/Na DIBs4.2675/200195.4[54]
CuTCNQ-MOFLi+ DIBs3.7570/50157.0[226]
F4TCNQ/
LLZTO-PEO
Li+3.4077/5080.00[55]
CuTCNQNa+2.00–4.1084/50255.0[202]
TCNQNa+2.20–3.0070/50233.0[239]
CuTCNQK+2.8075/50244.0[204]
TCNQ/CCPZn2+1.1078.54/1000171.0[186]
TCNQ/PPyZn2+0.4575.40/1000245.8[50]
p-TCNQZn2+1.1055.20/50250.0[189]
TCNQZn2+1.0298.90/50244.4[190]
FTCNQZn2+1.2093.40/50168.5[190]
F4TCNQZn2+1.4075.30/50135.1[190]
TCNQ/PNZZn2+1.00–3.8088/100150.0[193]
TCNQAl3+1.6075/2000180.0[245]
TCNQ/PPyAl3+0.45–0.8576/100245.8[43]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Starodub, T. Tetracyanoquinodimethane and Its Derivatives as Promising Sustainable Materials for Clean Energy Storage and Conversion Technologies: A Review. Sustainability 2025, 17, 10612. https://doi.org/10.3390/su172310612

AMA Style

Starodub T. Tetracyanoquinodimethane and Its Derivatives as Promising Sustainable Materials for Clean Energy Storage and Conversion Technologies: A Review. Sustainability. 2025; 17(23):10612. https://doi.org/10.3390/su172310612

Chicago/Turabian Style

Starodub, Tetiana. 2025. "Tetracyanoquinodimethane and Its Derivatives as Promising Sustainable Materials for Clean Energy Storage and Conversion Technologies: A Review" Sustainability 17, no. 23: 10612. https://doi.org/10.3390/su172310612

APA Style

Starodub, T. (2025). Tetracyanoquinodimethane and Its Derivatives as Promising Sustainable Materials for Clean Energy Storage and Conversion Technologies: A Review. Sustainability, 17(23), 10612. https://doi.org/10.3390/su172310612

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

Article Metrics

Back to TopTop