Next Article in Journal
Preparation of eGaIn NDs/TPU Composites for X-ray Radiation Shielding Based on Electrostatic Spinning Technology
Next Article in Special Issue
Advances in Nanoarchitectonics: A Review of “Static” and “Dynamic” Particle Assembly Methods
Previous Article in Journal
The Microstructural Evolution and Corrosion Behavior of Zn-Mg Alloys and Hybrids Processed Using High-Pressure Torsion
Previous Article in Special Issue
Gas-Sensing Properties and Mechanisms of 3D Networks Composed of ZnO Tetrapod Micro-Nano Structures at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Materials Nanoarchitectonics at Dynamic Interfaces: Structure Formation and Functional Manipulation

by
Katsuhiko Ariga
1,2
1
Research Center for Materials Nanoarchitectonics (MANA), National Institute for Materials Science (NIMS), 1-1 Namiki, Tsukuba 305-0044, Ibaraki, Japan
2
Graduate School of Frontier Sciences, The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa 277-8561, Chiba, Japan
Materials 2024, 17(1), 271; https://doi.org/10.3390/ma17010271
Submission received: 13 December 2023 / Revised: 25 December 2023 / Accepted: 29 December 2023 / Published: 4 January 2024
(This article belongs to the Special Issue Nanoarchitectonics in Materials Science)

Abstract

:
The next step in nanotechnology is to establish a methodology to assemble new functional materials based on the knowledge of nanotechnology. This task is undertaken by nanoarchitectonics. In nanoarchitectonics, we architect functional material systems from nanounits such as atoms, molecules, and nanomaterials. In terms of the hierarchy of the structure and the harmonization of the function, the material created by nanoarchitectonics has similar characteristics to the organization of the functional structure in biosystems. Looking at actual biofunctional systems, dynamic properties and interfacial environments are key. In other words, nanoarchitectonics at dynamic interfaces is important for the production of bio-like highly functional materials systems. In this review paper, nanoarchitectonics at dynamic interfaces will be discussed, looking at recent typical examples. In particular, the basic topics of “molecular manipulation, arrangement, and assembly” and “material production” will be discussed in the first two sections. Then, in the following section, “fullerene assembly: from zero-dimensional unit to advanced materials”, we will discuss how various functional structures can be created from the very basic nanounit, the fullerene. The above examples demonstrate the versatile possibilities of architectonics at dynamic interfaces. In the last section, these tendencies will be summarized, and future directions will be discussed.

1. Introduction

In considering the development of functional materials, it is significant to learn how a biological system creates a functional system. This is because in a biofunctional system, multiple components are rationally arranged, and they selectively perform their functions with a very high efficiency [1,2,3]. Each function is very dynamic, and several functions work in harmony. Such a high level of functional expression remains difficult to achieve in artificial materials systems. Many functions in biological systems are achieved by the accumulation of functional molecules at interfaces such as cell membranes [4,5,6,7]. Intermolecular interactions work together at these interfaces, and information such as electrons and energy flow in specific directions. Biological systems have established these superior mechanisms over billions of years of evolution. Most of these events are conducted at dynamic interfaces. The construction of material systems at dynamic interfaces will be the key for humans to imitate them in a short time and to develop advanced functional systems.
On the other hand, humanity has developed functional materials in the course of various scientific developments. The history of the development of functional materials corresponds to the history of human progress. Even today, the development of various functional materials is still necessary to meet the demands of energy [8,9,10,11,12,13,14,15,16,17], the environment [18,19,20,21,22,23,24,25,26,27], and biomedical requirements [28,29,30,31,32,33,34,35,36,37]. Modern science and technology must develop materials for various functions, such as for energy production [38,39,40,41,42,43,44,45], energy storage [46,47,48,49,50,51,52], separation [53,54,55,56,57,58,59,60], sensing [61,62,63,64,65,66,67,68], and drug delivery [69,70,71,72,73,74,75,76]. In the 20th century, academic areas for the creation of various materials have been established, and it has become possible to create materials on demand that do not exist in nature. This is supported mainly by materials-related chemistry. Humans have developed various types of chemistry since the 20th century, which has led to modern scientific research. Research on the development of functional systems through organic chemistry [77,78,79,80,81,82,83,84,85,86], inorganic chemistry [87,88,89,90,91,92,93,94,95,96], coordination chemistry [97,98,99,100,101,102,103,104,105,106], supramolecular chemistry [107,108,109,110,111,112,113,114,115,116], polymer chemistry [117,118,119,120,121,122,123,124,125,126], interface chemistry [127,128,129,130,131,132,133,134,135,136], various material chemistry [137,138,139,140,141,142], and bio-related chemistry [143,144,145,146,147,148,149,150,151,152] has continued unceasingly.
The science of functional material development and structural control, such as the creation of hybrids and composites [153,154,155,156,157,158], has raised awareness of the importance of not only the material itself, but also its internal nanostructure [159,160,161]. The turning point was the creation of the concept of nanotechnology. Nanotechnology has been advancing science and technology with the symbolic size of the nanoscale. It is continuously active today. We observe and manipulate objects at the atomic, molecular, and nanoscale levels [162,163,164,165,166,167]. Furthermore, we can analyze and understand phenomena in those nanoregions [168,169,170,171]. Such technologies have been applied to analytical chemistry [172,173,174,175,176] and physical chemistry [177,178,179,180,181], and the impact of the structure of a material on its function has been studied.
The next step in nanotechnology is to establish a methodology to assemble new functional materials based on the knowledge of nanotechnology. This task is taken by nanoarchitectonics [182,183,184,185]. Just as Richard Feynman founded nanotechnology in the 20th century [186,187], nanoarchitectonics was proposed by Masakazu Aono at the beginning of the 21st century [188,189]. Nanoarchitectonics can be considered as a post-nanotechnology concept [190]. This concept is not something completely new, but rather represents the integration of established fields: it combines nanotechnology with organic chemistry, inorganic chemistry, polymer chemistry, coordination chemistry, supramolecular chemistry, material chemistry, biochemistry, and microfabrication techniques (Figure 1) [191].
In nanoarchitectonics, we architect functional material systems from nanounits such as atoms, molecules, and nanomaterials. It architects functional materials through the selection and combination of atomic and molecular manipulations, physical or chemical transformations, self-assembly and self-organization, alignment by external fields and forces, microlevel and nanolevel fabrications, and biochemical processes [192]. With this method of combining several processes, hierarchical structures are more easily obtained [193] than in the process of self-assembly by simple equilibrium [194,195,196]. High functionality in biological systems is due to the flow of molecules, ions, electrons, and energy in hierarchical and asymmetric structures. Nanoarchitectonics is predisposed to construct such advanced concepts in biological systems [197]. In addition, nanolevel phenomena are prone to contain uncertainties. They may not be uniquely determined by thermal fluctuations, stochastic distributions, quantum effects, etc. [198]. The effects tend to be harmonized rather than summarized [199]. Nanoarchitectonics, which is founded on nanophenomena, has the characteristic of harmonizing effects. It also has similar characteristics to biofunctional systems, where the effects are dynamically harmonized in thermal fluctuations. In terms of the hierarchy of the structure and the harmonization of the function, material creation by nanoarchitectonics has similar characteristics to the organization of the functional structure in biosystems [200].
In biofunctional systems, a wide variety of functional molecules are organized. In order to mimic such systems, it is essential to have a material-creating capability for every object. In principle, nanoarchitectonics can be applied to any material without any limitations. All matter is composed of atoms and molecules. Nanoarchitectonics, which builds materials from molecules and atoms, is a methodology that can be applied to the creation of all materials. It can be a Method for Everything in material chemistry [201,202], analogous to the Theory of Everything in physics [203]. In fact, nanoarchitectonics has a wide range of applications. The research papers that claim to be nanoarchitectonics are as follows. The application of nanoarchitectonics can be seen in basic fields, such as the synthesis of functional materials [204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219], the control of specific structures [220,221,222,223,224,225,226,227,228,229,230], the creation of structures [231,232,233,234,235,236,237,238,239,240], basic physics [241,242,243,244,245,246,247,248,249,250,251,252], relatively basic biochemistry [253,254,255,256,257,258,259,260], and the development of simple functional materials for biological systems [261,262,263,264,265,266,267,268,269,270,271,272], as well as application-oriented fields such as catalyst development [273,274,275,276,277,278,279,280,281,282], sensors [283,284,285,286,287,288,289,290,291], devices [292,293,294,295,296,297,298,299], energy production [300,301,302,303,304,305,306,307], energy storage [308,309,310,311,312,313,314,315,316,317], the environment [318,319,320,321,322,323,324,325,326,327], biomedical [328,329,330,331,332,333,334,335,336,337,338,339,340,341], etc. This can be regarded as a list of some of the most important applications of the current technology.
The characteristics of nanoarchitectonics, such as hierarchical structure creation, function harmonization, and deployment in a variety of materials, are suited to the architecture of bio-like functional structures. Looking at actual biofunctional systems, as mentioned above, the dynamic properties and interfacial environments are important keys. In other words, nanoarchitectonics at dynamic interfaces is important. In particular, thin film fabrication at the interface will be an important methodology for this target [342,343,344]. One method for forming dynamic structures at interfaces that do not move, such as individuals, is the vacuum deposition method [345,346,347]. Thin film nanoarchitectonics has also been used as a thin film creation material, such as with DNA-related materials that exhibit dynamic properties [348]. Layer-by-layer (LbL) assembly is a well-known methodology for the nanoarchitectonics of thin films on solid substrates or colloidal particles in a wet process [349,350,351,352,353]. In this technique, a very wide range of materials are sequentially stacked in the desired manner with a simple operation. The interfaces where adsorption occurs are mostly solid substrates or solid colloids, which do not move dynamically. However, the thin films that are stacked are mostly amorphous and have dynamic motion and function [354,355,356]. A typical example of a thin film preparation method that uses a dynamic interface with free motion is the Langmuir–Blodgett (LB) method, which uses a gas–liquid interface [357,358,359]. Thin films deployed on a liquid interface (mainly a water surface) are freely compressed, aligned, controlled, and organized. The thin films can be transferred to a solid substrate as a multilayer structure [360,361,362]. In recent years, the vortex LB method, which promotes orientation by introducing a vortex flow at the air–water interface, has also been devised [363]. Ultra-high temperature LB methods have also been developed that use conditions of 100 °C [364] or 200 °C [365], rather than the usual room temperature conditions. Organizing methods using not only the air–water interface but also the liquid–liquid interface have also been reported [366,367]. These methods can be called nanoarchitectonics methods using dynamic interfaces.
Nanoarchitectonics is the concept of architecting materials from nanounits. The ultimate model for reference is the highly organized functional structure of biological systems. The big challenge of nanoarchitectonics is related to the establishment of a unified concept that produces bio-like functional materials systems from essential atoms, molecules, and nanounits, rather than accomplishments in particular fields. The key to this is the interfacial environment and its dynamic nature. In this review paper, nanoarchitectonics at dynamic interfaces will be discussed, looking at typical recent examples. In particular, the more basic topics of “molecular manipulation, arrangement, and assembly” and “material production” will be discussed in the first two sections. Then, in the following section, “fullerene assembly: from zero-dimensional unit to advanced materials”, we will discuss the creation of various functional structures from the very basic nanounit, the fullerene. These examples will demonstrate the versatile possibilities of architectonics at dynamic interfaces. In the last section, these trends will be summarized, and future directions will be discussed.

2. Molecular-Level Dynamic Process: Manipulation, Arrangement, and Assembly

Compared to crystals and solid surfaces, the degree of freedom of molecular motion is much higher in liquid systems. In a solution system spread over three dimensions, the degree of freedom of molecular motion is too large to control. Unlike three-dimensional systems, liquid interfaces such as gas–liquid and liquid–liquid are used as externally controllable fields that preserve the degree of freedom of molecular motion by the liquid. In particular, the Langmuir–Blodgett method has been used to control molecules at the gas–water interface in various ways.
Negi et al. synthesized some variously designed chiral molecules and analyzed their behavior on the water surface to control their molecular orientation and to verify the molecular structures that are suitable for supramolecular structures and nanostructures [368]. The compound consists of a pyrene ring introduced into the asymmetric carbon portion of a secondary alcohol, which is the hydrophilic portion of an amphiphilic molecule with a pyrene sulfonic acid group. As concrete molecules, 1-stearylpyrene-6-sulfonate and 1-stearylpyrene-8-sulfonate were synthesized, and their monolayer behavior was analyzed. In addition to the conventional method of π-A isotherms, in situ surface fluorescence spectroscopy, Brewster angle microscopy, and atomic force microscopy were used to study the assembling behavior. In particular, the structures of the racemic and optically active monolayers were analyzed. The monolayers that formed at the air–water interface of the racemic compounds resembled a racemic solid solution. Furthermore, the optically active compound formed a monolayer with a well-controlled structure. This is due to the cooperative action of the face-to-face association of the pyrene rings and chiral steric factors. The combination of π-π stacking and chirality allows for some control of the monolayer structure. In order to control the structure of the molecular assemblies by dynamic movements at the interface, it is important to use chirality more effectively. For this purpose, further improvement of the molecular design is needed. To control the molecular assemblies more precisely, it is also important to combine chirality with other intermolecular interactions, a combination that is useful for the design of supramolecular structures and nanostructures.
Negi et al. also investigated the pressure dependence of the aggregation nanoarchitectonics of binaphthyl-type amphiphiles at dynamic interfaces [369]. In this study, 2, 2′-bis(octadecyloxy)-1, 1′-binaphthyl-6, 6′-dicarboxylic acid was designed as a monolayer-forming molecule. This molecule has an axial chirality derived from the binaphthyl moiety with two COOH groups. The monolayer behavior of the racemic and optically active S forms was investigated. Brewster angle microscopy, atomic force microscopy, and surface pressure–area isotherms were used to analyze the monolayer structure. The racemic form creates a solid film, while the S-isomer forms a liquid film. It was also found that the real lattices of these monolayers differ significantly (Figure 2). The differences in the real lattice are due to steric regularity resulting from axial chirality, which causes differences in the mode of intermolecular interactions. In the racemic case, the solid membrane is formed by a cooperative network of π-π interactions between the naphthyl rings and of strong van der Waals interactions. This model is consistent with the actual lattice obtained from atomic force microscopic images. In the S-isomer, the intermolecular distance is greater than in the racemic form, so strong intermolecular forces are less likely to act. Relatively weak interactions are expected to result in a one-dimensional arrangement of long alkyl chains and naphthyl rings. At a low surface pressure, the van der Waals interactions between the monolayers are small and the intermolecular forces are relatively weak. However, upon compression, a phase transition from liquid-expanded film to liquid-condensed film was observed at the π-A isotherm. It is considered that the two-dimensional physical compression at the dynamic interface reduced the dihedral angle of the binaphthyl ring in the one-dimensional columnar structure, narrowing the intermolecular distance and thus strengthening the intermolecular interactions, which induced the phase transition.
Takase et al. focused on a highly conductive charge transfer complex of (phthalocyaninato)cobalt iodide [370]. They investigated the effect of the enhanced conductivity of cobalt phthalocyanine crystals on catalytic activity. This is also beneficial for the industrial application of cobalt phthalocyanine as a catalyst for CO2 reduction. They applied the cobalt phthalocyanine crystal phase transition method to develop a useful method for synthesizing (phthalocyaninato)cobalt iodide by simply mixing a KI solution containing CF3COOH and cobalt phthalocyanine with a CH2Cl2 solution at the interface (Figure 3). UV-vis absorption spectra showed that under acidic conditions, the iodide ion (I) changes to triiodide ion (I3) at the interface between the aqueous potassium iodide solution and the organic solvent. In other words, when the aqueous KI solution was mixed with an organic acid solution containing cobalt phthalocyanine, the I changed to I3, forming a highly conductive charge transfer complex of (phthalocyaninato)cobalt iodide upon cobalt phthalocyanine with the I3 stacked separately. The catalytic properties of (phthalocyaninato)cobalt iodide were investigated with polarization measurements and electrochemical impedance spectroscopy using a gas diffusion carbon electrode. The (phthalocyaninato)cobalt iodide nanoarchitectonized at the dynamic interface and showed high catalytic activity for CO2 reduction and high CO formation selectivity. It is expected that such attempts can be applied to electrochemical devices.
Dynamic interfaces can also be a field for molecular manipulation [371,372]. For example, the coupling of macroscopic mechanical manipulations and molecular functions, which are phenomena that differ greatly in size, is possible in a field with reduced dimensions, such as in a field like a dynamic interface [373,374]. In an asymmetric interfacial two-dimensional system such as a molecular thin film, the film has a macroscopic dimension in the in-plane direction and a molecular dimension in the thickness direction. Mechanical macroscopic motions in the in-plane direction of the monolayer may be reflected in the molecular function. By compressing the monolayer by tens of centimeters, nanometer-sized molecular machines and structures can be easily controlled. For example, Adachi et al. have developed a new principle called “submarine luminescence” as a luminescence control, involving the molecular manipulation of double-paddle platinum complexes at the air–water interface (Figure 4) [375]. When a monolayer of the double-paddle Pt complexes is formed at the air–water interface and the water surface is compressed from both ends, the luminescence intensity increases rapidly from a certain compression state. The double-paddle platinum complexes are used to form square planar complexes. These double-paddle platinum complexes exhibit unique molecular activities such as flapping and the relative rotation of the double plane, and the luminescent sites of the double-paddle platinum complexes float from the aqueous phase, a high dielectric medium, to the gas phase, a low dielectric medium. The luminescence intensity increases as the luminescent sites levitate from the aqueous phase. Spectroscopic measurements and DFT calculations confirm that the molecular arrangement pattern changes simultaneously with the luminescence switching. The luminescence ability of the monolayer is determined by whether the luminescent sites are present in water or air. At a low surface pressure, the H-shaped complex of the double-paddle platinum complex sinks in water and has a weak emission due to hydrogen bonding. The double-paddle platinum complex is pushed to the surface by mechanical compression, so that the two planes behave independently at the air–water interface. One face in the aqueous phase is suppressed, while the other face in the gas phase is freed from the excitation energy dispersion due to the molecular contact. This is an example of using the difference between two adjacent environments, the aqueous and gas phases, to control physical properties. This is a “submarine luminescence” mechanism, in which the optical properties are controlled in a submarine-like manner through molecular manipulation in a nanometer-sized asymmetric interfacial environment. This opens up new possibilities for molecular manipulation using the dynamic behavior of the air–water interface.
Maeda et al. reported a method that allows flexible control of the intensity and signature of circularly polarized emissions from molecular aggregates at a dynamic interface with rotational motion (Figure 5) [376]. In this study, an achiral trans-bis(salicylaldiminato)platinum(II) complex is deployed at the air–water interface. The circularly polarized luminescence of the aggregates was precisely controlled by creating clockwise and counterclockwise vortexes at the air–water interface and by specifically aggregating them at various motion velocities. The two-dimensional aggregates formed at the air–water interface under the vortex flow are quite different from the three-dimensional molecular arrangement of crystals formed from organic solutions. The helical torsional forces of the vortex flow result in the supramolecular chirality of the two-dimensional domains caused by the one-way torsion in the coordination plane stacking. The sense of torsion is determined by the direction of the vortex rotation, and the degree of torsion is controlled by the vortex velocity. The complete and infinite stacking of trans-bis(salicylaldiminato)platinum(II) coordination planes induces dispersion of the optical energy of the excited state. The platinum coordination planes remain in mutual contact with each other in a continuous, shallow stacking interaction during the chiral transition via helical twisting at the organic–water interface. The enhancement of luminescence by eddy current control is attributed to the change in the stacking status of the coordination planes depending on the vortex velocity. This method of nanoarchitectonics at dynamic interfaces is a milestone example of precisely controlled, circularly polarized emission from molecular aggregates.
Emrick, Russell, and co-workers examined the aggregation behavior of bottlebrush polymers at dynamic interfaces (Figure 6) [377]. The aim of this study was to elucidate the fundamental relationships between the nanostructural and macroscopic properties of various bottle brush polymers. The results answer the question of how the properties of all fluid-printed structures can be controlled by molecular design and their properties. Bottle brush polymer surfactants are formed by the interfacial interaction between a bottle brush polymer with poly(acrylic acid) side chains dissolved in the aqueous phase and an amine functionalized ligand dissolved in the oil phase. The bottlebrush polymer surfactants were formed at the water–oil interface through the electrostatic interaction of carboxylate (poly(acrylic acid)) and ammonium (aminopropylisobutyl polyhedral oligomeric silsesquioxane). The bottlebrush polymer surfactant strongly associates and binds to the liquid–liquid interface. The initial aggregation rate, interface filling efficiency, and stress relaxation are specified by the ratio of the degree of polymerization of the backbone to that of the side chains. Bending stiffness as a mechanical property and stress relaxation behavior during compression of the aggregates were also studied. The ratio of the degree of polymerization of the polymerization side chains of the backbone changed the effective area projected on the fluid interface of the polymer from spherical to cylindrical. The results provide general design guidelines for the printing of structured liquids with bottle brush polymers. In addition, the high density of functional groups at the printed liquid interface of the bottlebrush polymer interfacial activity allows for further hierarchical nanoarchitectonics, such as incorporating metal cations and active enzymes. These structures could also be applied to applications such as two-phase reactors where interfacial chemistry can be controlled.
A dynamic interface provides a field in which the degree of freedom of motion is guaranteed to some extent and does not diverge significantly. In such a field, it is possible to manipulate molecules and create unique molecular assemblies by external forces. For example, external forces can manipulate the molecular position between the two phases of air and water to control the luminescence behavior. The direction and intensity of the vortex flow at the interface can tune the circularly polarized luminescence of the assembly mode of the functional complex molecules. It is expected that the successful combination of appropriate molecular design and dynamic interfacial nanoarchitectonics will enable a wide variety of molecular manipulation and assembly control.

3. Material Production: Dynamic Materials-Level Formation Process at Interfaces

At dynamic interfaces, it is possible to manipulate molecules and control their aggregate morphologies. When those molecules react or combine, they become materials. This also happens with inorganic materials and their hybrids. Dynamic interfaces provide an appropriate site for material production, not just limited to molecular assembly.
Dynamic interfaces, such as the liquid–liquid interface, provide a very good venue for creating biomimetic structures such as artificial cell membranes [378,379,380]. One of the candidates for the creation of biomimetic structures is cell membrane models. In such models, permeability controls through cell-mimic membranes are attractive research targets. Patra and co-workers used dynamic imine chemistry at the oil–water interface to control mass permeability (Figure 7) [381]. The effect of dynamic covalent bonding was controlled through modulation of the droplet shape at the dynamic interface. Imine bond formation between water-soluble polyethyleneimine and oil-soluble aromatic aldehydes significantly reduces the interfacial tension and greatly stabilizes the oil–water interface. Furthermore, the interfacial tension can be tuned through varying the degree of aldehyde (mono-, di-, and tri-) functionality in the imine bond formation. When a compressive force is applied to the droplet, the imine-mediated aggregation is successfully packed. Anisotropic compartmentalization of the liquid–liquid interface thus occurs. The pH dependence of the Schiff base reaction is skillfully exploited to shift the equilibrium, leading to reversible toggling from jamming to unjamming in the interfacial assembly. The stimuli-responsive behavior of the dynamic imine assembly was demonstrated in a pH-triggered control of content release through the interfacial membrane. At a pH of 10, maximal cross-linking through the imine bonds occurred at the interface. As a result, the diffusion of the dye molecules through this interfacial membrane was greatly inhibited (the locked state). On the other hand, at a pH of 2, the opposite condition occurred (the unlocked state), where the interfacial cross-linking was negligible and the dye release was maximal. These results are expected to contribute significantly to the field of dynamic covalent assembly at the liquid–liquid interface, droplet structuring, and various types of stimuli-responsive membranes.
Dynamic imine chemistry, which deals with the Schiff base formation, is one promising approach for the synthesis of well-defined molecular structures. In imine chemistry, the dynamic imine bond formation properties allow for a self-healing mechanism. Thus, molecular structures can be formed with thermodynamic control. Widely applied to small molecule and polymer synthesis, it is useful for the synthesis of two-dimensional polymers at interfaces. Zhang and co-workers synthesized a covalent monolayer of two-dimensional polymer under ambient temperature conditions at the air–water interface (Figure 8) [382]. The prepared two-dimensional polymer sheets, which were transferred from a Langmuir trough via a horizontal Schaeffer-type transfer. The thickness of the sheet (0.7 nm), determined by atomic force microscopy (AFM), corresponded to that for its monolayer. A freestanding monolayer sheet was obtained, with lateral dimensions in the range of tens of micrometers. Within the resolution range, the presence of imine bonds and the absence of terminal groups were indicated through the Raman spectra. Smooth, coherent, large, freestanding polyimine monolayers were nanoarchitectonized. Such nanoarchitectonics of conjugated two-dimensional polymers at dynamic interfaces will contribute to providing materials for electronic and optoelectronic applications.
As described above, material synthesis at the dynamic interface, the air–liquid interface, is a promising method for obtaining nanosheets because the components grow spontaneously in a two-dimensional direction. Various types of nanosheets have been synthesized at dynamic interfaces from various functional molecules. Metal–organic framework (MOF) and covalent organic framework (COF) nanosheets are nice examples [383,384,385,386,387]. MOF nanosheets have attracted significant attention as components of electronic devices such as electrocatalysts and chemical resistivity sensors. However, achieving the desired properties with reasonable control over size and shape is not always easy. Makiura and co-workers investigated the effect of the type of solvent in the ligand diffusion solution to control the formation of the MOF nanosheets (Figure 9) [388]. In this MOF nanosheet synthesis, the ligand 2,3,6,7,10,11-hexaiminotriphenylene was spread on an aqueous solution containing Ni2+. The influence of the solvents used, methanol and N,N-dimethylformamide, was investigated. The usual role of solvents in the Langmuir film formation is to disperse molecules at the air–liquid interface without causing aggregation. In addition to this, the dynamic synthesis of the MOF nanosheets at the air–water interface also involves effects on coordination bonds and p-p interactions. Different expansion solvents alter the uniformity, morphology, lateral domain size, and chemical composition of the nanosheets. The macroscopic morphological uniformity of the MOF nanosheets is higher when N,N-dimethylformamide is used as the expansion solvent than when methanol is used as the solvent. The evaporation of the highly volatile methanol results in a variety of interfacial states, including surface ripping and solvent subphase miscibility. This results in variations in morphology on the micrometer scale. Conversely, N,N-dimethylformamide is less volatile and easily binds to the amino groups of the ligand. This facilitates the formation of coordination bonds and the growth of nanosheets. These insights into MOF nanosheet synthesis at dynamic interfaces provide the fundamental knowledge needed to further optimize interfacial synthesis methods. Full control of the nanoarchitectonics of MOF nanosheets will enable the prediction of chemical and physical properties closely related to the properties of the nanosheets. This will accelerate their use in diverse nanodevices.
The dynamic interface is an ideal place to synthesize two-dimensional materials. Carbon materials such as carbon nanosheets, represented by graphene, are attracting attention as novel electronic, optical, and catalytic materials [389,390,391]. In addition to conventional methods, the bottom–up method of carbon nanosheet synthesis, in which molecules are assembled into carbon through self-assembly, is attracting attention. The advantage of this bottom–up method is that the structure formed by the molecules can be precisely controlled at the nanolevel. However, the bottom–up method requires knowledge of advanced supramolecular chemistry. Therefore, a method to obtain nanosheets using a simple technique that anyone can do is being sought. Mori et al. reported the fabrication of mesoporous thin films with both uniformity and many tens of nanometer pores through a simple method in which a vortex flow is created in a beaker filled with water and carbon nanorings, which are ring-shaped carbon molecules that float on the surface of the water and are then transferred to a substrate (Figure 10) [392]. A method that can create a vortex flow at the water surface and orientate and accumulate materials with that flow is called the vortex Langmuir–Blodgett method (vortex LB method). This is a nanoarchitectonics method that utilizes dynamic material movement at the interface. The resulting thin films were carbonized under an inert gas atmosphere to synthesize carbon nanosheets. Interestingly, this molecular thin film composed of carbon nanorings had numerous pores (mesoporous) of several tens of nanometers. Even after carbonization through calcination, the carbon nanosheets retained this mesoporous structure. Before calcination, the nanosheet is an insulator that does not conduct electricity. However, after calcination, the carbon nanosheets have been transformed into conductors. In other words, the carbon is bonded together through calcination to form carbon nanosheets with a strong network. The same process in the presence of pyridine and carbon nanorings produced N-doped carbon nanosheets with an unexpectedly high nitrogen content. X-ray photoelectron spectroscopy (XPS) showed that the nitrogen in these carbon nanosheets has an electronic state that exhibits useful catalytic activity. The thin film fabrication method used in this study requires only a very small amount of carbon nanorings (1 ng) to fabricate a 1 m2 nanosheet. Furthermore, the technology can be industrially deployed by increasing the area of the nanosheet to a large area. In particular, the synthesized N-doped carbon nanosheets are expected to be used for various applications, such as efficient catalysts for oxygen reduction reactions in high-performance fuel cells and high-performance electrochemical supercapacitors. In addition, their mesoporous structure with a large surface area is expected to be applied to fuel cells and other applications as a catalyst that can replace expensive platinum.
Low temperature solution processes for thin film semiconductors are more cost-effective than conventional vacuum processes. However, they lead to more defects and residual tensile stress during high-speed bulk crystallization. Chen, You, Liu, and co-workers have developed a new strategy called dynamic liquid crystal transition using interfaces (Figure 11) [393]. The above problem can be solved in one step. The new strategy, dynamic liquid crystal transition, is designed to spontaneously delay bulk crystallization and release residual strain at the interface in situ during annealing. In the design principle of perovskite thin films using this method, the dynamic liquid crystal transition molecules first interact with the bulk perovskite crystal grains, while spontaneously healing the interface via a dynamic transition. Thermotropic liquid crystal molecules are used to demonstrate this strategy. Liquid crystal molecules interacting with perovskite colloids form intermediate adducts that delay the crystallization. The interaction of lead iodide with the target molecules in the solution delays the crystallization process and increases the crystal grain size. During the dynamic transition of the target molecules, they leave the perovskite and attach to the carrier transport layer. Target molecules concentrated at the interface tune the thermal mismatch. They act as a buffer to release residual strain between the perovskite and the carrier transport layer. This technique also significantly improves the environmental stability and photostability of the perovskite. These techniques envision a new engineering method of additive phase transitions for high-performance perovskite solar cells. The dynamic liquid crystal transition strategy can be a general method to achieve both delayed crystallization and spontaneous lower interface healing in a single step in the solution processing of thin film semiconductors.
Various batteries, such as lithium-ion batteries [394,395,396,397,398] and potassium-ion batteries [399,400,401], have attracted significant attention. Zinc-ion batteries have the advantage of high safety and have also attracted attention from those focused on environmental protection [402,403,404,405]. However, there is the critical problem of dendrite formation and the rapid decrease in battery life due to Zn debris that accumulates on the surface. To resolve this problem, Fan, Zhang, and co-workers developed a self-adaptive polydimethylsiloxane/TiO2-x coating that can dynamically adapt to volume changes and inhibit dendrite growth (Figure 12) [406]. The polymer synthesized here, polydimethylsiloxane, has high dynamic adaptability due to the micro-crosslinking of B-O bonds. There is sufficient time for the B-O bonds to break, and entanglement of the molecular chains prevents molecular deformation. Therefore, polydimethylsiloxane is fluid on the macroscale. With an increasing strain rate, the characteristic breaking time of the B-O crosslink becomes dominant. The B-O bonds resist the movement of the molecular chains and exhibit some degree of elasticity. The combination of TiO2−x with high oxygen vacancies and polydimethylsiloxane induces rapid and uniform migration of Zn2+. The coated anode is characterized by excellent cycling stability and shows great potential as an anode for Zn ion batteries. This nanoarchitectonics of dynamically adapting artificial coatings may be a beneficial method for the protection of various electrodes.
As seen in molecular manipulation and molecular assembly, anisotropic assemblies of materials are formed at dynamic interfaces. Under such conditions, reactions such as covalent bonding can produce ultrathin film forms with unique structures such as ordered porous materials including metal–organic frameworks (MOFs), covalent organic frameworks (COFs), and nanocarbon ultrathin films. Dynamic interfaces are an ideal place for the production of nanocarbon ultrathin films. At dynamic interfaces, reactants from both phases that form the interface dynamically come together to form the material. There is tremendous potential for the selection of components. The products often remain at the interface and can be easily recovered. Many possibilities remain in this area, and it is expected to develop not only in academic research but also in industrial applications.

4. Fullerene Assembly: Examples of Dynamic Formation Process from Zero-Dimensional Unit to Advanced Materials

An interesting example of nanoarchitectonics at dynamic interfaces is the formation of supramolecular assemblies at the liquid–liquid interface of fullerenes (C60, C70, etc.). When a fullerene is dissolved in a well-soluble solvent and a fullerene-poor solvent is added, a variety of supramolecular assemblies and crystals are formed at the liquid–liquid interface [407]. Fullerenes are composed of a single element, carbon, and have a simple zero-dimensional shape. Nevertheless, by simply changing the conditions, one-dimensional [408,409], two-dimensional [410,411], three-dimensional [412,413], and hierarchical structures [414] can be obtained. Those structures can also be treated at high temperatures to synthesize various forms of carbon materials. These systems reflect the diversity of nanoarchitectonics at dynamic interfaces. In the following, some recent examples are given.
Hollow carbon spheres are widely expected to be used in a variety of technological fields, including energy conversion and storage, catalysis, adsorption, drug delivery, and nanodevices [415,416,417]. However, it is not always easy to control the properties and morphology of the resulting hollow carbon materials. Chen et al. have developed kinetically controlled liquid–liquid interfacial precipitation, in which solid or hollow C60 nanospheres are kinetically controlled liquid–liquid interfacial precipitation that can control morphology and size [418]. This strategy is called the kinetically controlled liquid–liquid interfacial precipitation (KC-LLIP) strategy (Figure 13), where ethylenediamine was selected as a covalent cross-linker of C60, and the formation of C60-ethylenediamine products was controlled by the addition of isopropyl alcohol. To introduce the hollow structure, ethylenediamine-sulfur was used as an in situ generated droplet to form the yoke–shell structure. The design of the hollow C60 spherical structures was based on controlling the growth rate of the ethylenediamine-sulfur yoke and the C60-ethylenediamine shell. Thus, porous spheres, string hollow spheres, hollow spheres, and aperture hollow spheres were obtained. Some examples are as follows. Immediate deposition of C60-ethylenediamine on the surface of the first dispersed ethylenediamine-sulfur droplet yields porous spheres. The ethylenediamine-sulfur droplet is successfully incorporated as a template, so that the interior of the porous sphere contains many pores. The formation of string-like hollow spheres occurs by adding an appropriate amount of isopropyl alcohol to the ethylenediamine solution to reduce the growth rate of C60-ethylenediamine and promote the diffusion of ethylenediamine-sulfur droplets in the m-xylene solution. Hollow spheres were obtained when ethylenediamine-sulfur droplets coalesced into larger spheres during C60-ethylenediamine aggregation. Excess isopropyl alcohol causes the rapid formation of deformed ethylenediamine-sulfur droplets, resulting in the formation of sealed deformed spheres. The formation of open-aperture hollow spheres is due to the inability of the C60-ethylenediamine nucleus to precipitate into the shell aperture if it is occupied by other ethylenediamine-sulfur droplets during the coalescence. These diverse nanoarchitectonics represent an effective strategy for synthesizing C60 spheres, whose dimensions and morphology can be tuned through various kinetic controls using droplets as templates. Moreover, this nanoarchitectonics approach is characterized by its ease of implementation on a large scale. The resulting C60 nanospheres can be applied as effective carbon materials for advanced applications.
Certain amine reagents, such as ethylenediamine, are active reactants of fullerenes. Based on this activity, there is a technique for nanoarchitectonics of aggregates while reacting in situ. This technique has a high potential for the preparation of highly integrated structures and nitrogen-doped materials [419]. Chen et al. applied the in situ reaction method to the self-assembly process of C60 molecules and melamine/ethylenediamine components in solution (Figure 14) [420]. As a result, they succeeded in producing a new fullerene assembly, a micron-sized, two-dimensional, amorphous-shaped, regular object, a fullerene rosette. The two-dimensional fullerene rosette consists of six petals partially resembling hexagonal sheets, with a very smooth surface with thicknesses in the 120–165 nm range. No clear crystalline features in the two-dimensional plane were detected from the XRD patterns. The two-dimensional fullerene rosettes were immobilized in a quartz crystal microbalance to create sensors for a variety of volatile organic compounds. The frequency shifted quickly when exposed to volatile organic compounds. When the solvent vapor was removed from the chamber, it returned to almost its initial state. Thus, reversible vapor adsorption/desorption was suggested. The detection sensitivity for a wide range of volatile organic compounds obeys as follows, in order: formic acid > acetic acid > pyridine > acetone > aniline > benzene > ethanol > ethyl acetate > toluene > cyclohexane > hexane. Formic acid sensing was compared to other carbon materials, hexagonal nanosheets, fullerene nanotubes, acid-treated fullerene nanotubes, and commercial activated carbon. The two-dimensional fullerene rosette-modified quartz resonator electrode showed a remarkable performance compared to the other materials tested. The high ability to sense acid vapors such as formic and acetic acids is probably due to the abundance of amino groups in the aggregates. The rosette-based quartz crystal resonator sensor was also superior in the subtle size discrimination of formic and acetic acids. Another advantage of the developed fullerene rosette-based sensor is that it can be fabricated and measured at room temperature, avoiding a high temperature processing.
Wei at al. reported the vapor sensing performance of self-assembled, cone–husk fullerene C60 crystals prepared by dynamic liquid–liquid interface precipitation under an ambient temperature and pressure (Figure 15) [421]. Due to the unique semi-open tubular structure and typical micropore size, the cone–husk fullerene C60 crystals exhibit a highly sensitive detection performance for acetic acid vapor. The object uses a very quick liquid–liquid interfacial precipitation method. Isopropyl alcohol was quickly added to a fresh saturated solution of C60 in mesitylene and immediately shaken vigorously by hand for about 3 s. The cone–husk fullerene C60 crystal precipitate was washed with isopropyl alcohol to remove the organic solvent. They were separated from the mixture by centrifugation and dried in a vacuum oven at 70 °C for 3 h. Electron microscopic analysis confirmed that the cone–husk fullerene C60 crystals have a unique thin tubular wall with a semi-open structure at one end and a solid structure at the other end. Cone–husk fullerene C60 crystals have a microporous structure with crystalline pore walls. This structure provides the nanospace necessary for the adsorption of guest vapors. As a result, the gas-sensing performance is enhanced. The sensitivity of quartz crystal resonator sensor electrodes modified with cone–husk fullerene C60 crystals is in the following order: acetic acid > formic acid > ethanol > water > acetone > pyridine > formaldehyde > methanol > toluene > aniline > cyclohexane > hexane > 2-propanol. The diffusion of acetic acid vapor is facilitated by the unique thin tubular wall with a semi-open structure at one end of the cone–husk fullerene C60 crystal, which facilitates the diffusion of acetic acid vapor. This provides the nanospace necessary for the adsorption of acetic acid guest vapor and improves the gas-sensing performance. Cone–husk fullerene C60 crystals have great potential in the development of systems for volatile organic compound sensing that are selective for aliphatic acid vapors.
Tang et al. achieved a supramolecular assembly of fullerene C60 into high aspect ratio quasi-two-dimensional microbelts at room temperature at the dynamic liquid–liquid interface of a carbon disulfide solution of fullerene C60 and isopropyl alcohol (Figure 16) [422]. The fullerene microbelt was synthesized using the static liquid–liquid interfacial precipitation method. The length of the fullerene C60 microbelt can be controlled by the synthesis conditions. By optimizing the conditions, a fullerene microbelt with a length of 1 cm could be synthesized. The quasi-two-dimensional-structured fullerene microbelt was expected to have excellent mechanical stability, flexibility, transparency, and chemical stability. The fullerene microbelt can be converted to a mesoporous carbon microbelt with an amorphous or graphite backbone structure through carbonizing at 900 °C or 2000 °C, respectively. Nanoarchitectonics by high temperature heat treatment results in a novel bimodal porous carbon material derived from fullerene crystals, a crystalline p-electron rich carbon source. It exhibits higher specific capacitance than graphene, conventional mesoporous carbon, and other similar carbon materials, and it demonstrates an excellent electrochemical supercapacitor performance. The quasi-two-dimensional microbelt morphology with large pore volume is advantageous for enhanced charge transport in supercapacitor electrodes. They can be a potentially excellent material for energy storage applications.
Self-assembly and subsequent material creation at dynamic interfaces using fullerenes have been reported far beyond what is shown here. It is astonishing that fullerene, a single element, zero-dimensional object, a nanounit without anisotropy or specific functional groups, can create such a wide variety of materials. It shows the great potential of nanoarchitectonics, especially dynamic interfacial nanoarchitectonics.

5. Summary and Perspectives

In this review, nanoarchitectonics at dynamic interfaces as a method to create more functional materials has been discussed. Nanoarchitectonics is a methodology to architect functional materials using nanounits such as atoms, molecules, and nanomaterials. Since the method is applied without limiting the materials, it can be generally applied to almost all materials. A similar architecture and organization of functional structures is often seen in biological systems. In these systems, the interfacial environment and dynamic properties are strongly expressed. As a result of this background, therefore, it is suggested that nanoarchitectonics at dynamic interfaces is a promising method for the development of functional materials.
To this end, liquid interfaces provide a beneficial environment. Compared to crystals and solid surfaces, liquid interfaces have a much higher degree of freedom of motion for the molecules and materials contained therein. Furthermore, since the range of motion and dispersion is limited to a two-dimensional system, it is more controllable than a solution system spread over three dimensions. The general trends of the corresponding developments can be summarized with the listed examples. Liquid interfaces, such as gas–liquid and liquid–liquid interfaces, are important fields for nanoarchitectonics because these interfaces have a high degree of freedom in terms of their responsiveness to external stimuli. Their dynamic functions can also be controlled. Molecular manipulation and the formation of specific molecular aggregates are performed by external forces. In the example discussed above, luminescence behavior can be controlled through manipulating the molecular position between the two phases at the air–water interface. Alternatively, the direction and intensity of the vortex flow at the interface can tune the circularly polarized luminescence of the assembly style of the functional complex molecules. When such assembled molecules are reacted or combined, they become functional materials. In other words, the dynamic interface is a suitable field for material production. For example, it is an ideal site for the fabrication of metal–organic frameworks (MOF), covalent organic frameworks (COF), and nanocarbon ultrathin films. At dynamic interfaces, the reactants from both phases that form the interface dynamically come together to form materials. There is tremendous potential for the selection of components. Furthermore, it has been demonstrated that a wide variety of structures and materials can be created by the nanoarchitectonics of fullerenes, which are units composed of a single element, carbon, and having a simple zero-dimensional shape. As a method itself, the nanoarchitectonics of dynamic interfaces has the power to create a diverse group of materials. In addition to that, there is a high degree of freedom in the selection of components at the interface, with the result that it has tremendous potential. In addition to academic research, the development of industrial applications is also expected.
The nanoarchitectonics of dynamic interfaces has many possibilities for development. One of them is its application to artificial functional structures such as devices. In such fields, top–down processing technologies such as microfabrication have been applied [423,424,425]. This can be converted to a bottom–up approach through introducing nanoarchitectonics at dynamic interfaces. In this case, many functional molecules and their supramolecular structures will be used for device fabrication [426,427,428,429,430]. Recently, Ishii, Yamashita, and co-workers have developed a chemical doping technique based on the redox reaction of benzoquinone and hydroquinone in an aqueous solution under ambient conditions [431]. The doping levels varied with the pH of the aqueous solution, and the conductivity was accurately and consistently controlled over a wide range of about five orders of magnitude. This was performed at polymer organic semiconductor thin film interfaces but could also be applied to dynamic liquid interfaces. When such semiconductor control techniques are coupled with the concept of nanoarchitectonics at dynamic interfaces, bottom–up device production becomes a reality. The history of the device industry, which has relied on physical fabrication techniques, can be rewritten due to nanoarchitectonics.
Another possibility is the challenge of building material systems in which many types of functional molecules are rationally organized, as is the case in biological systems. Until now, supramolecular chemistry has produced an organization of a relatively small number of types of molecules. By combining them, it will be possible to build complex and hierarchical functional architectures. However, this is not an easy task. Each set of components requires its own conditions and molecular design. It is not easy to optimize a large set of such components. Fortunately, humanity has developed artificial intelligence and incorporated it into materials science. For example, machine learning has proven to be one of the solutions to the search for answers in materials and chemistry research [432,433,434,435,436]. The related concept of materials informatics has also been proposed [437,438,439]. In fact, there are proposals to integrate nanoarchitectonics and materials informatics in the development of functional porous materials and others [440,441]. The ultimate goal of nanoarchitectonics is to create complex and highly functional material systems such as those found in biological systems. As discussed in this review, the interfacial environment and dynamic behavior are very important factors. In addition, the incorporation of new methods such as materials informatics is essential for the further development of nanoarchitectonics in the future. One of the remaining challenges is the application of these strategies at larger industrial scales. Therefore, technological advancements must also be considered as important consequences.

Funding

This study was partially supported by JSPS KAKENHI, grant number JP20H00392 and JP23H05459.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Jordan, P.; Fromme, P.; Witt, H.; Klukas, O.; Saenger, W.; Krauß, N. Three-dimensional structure of cyanobacterial photosystem I at 2.5 Å resolution. Nature 2001, 411, 909–917. [Google Scholar] [CrossRef] [PubMed]
  2. Umena, Y.; Kawakami, K.; Shen, J.R.; Kamiya, N. Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 Å. Nature 2011, 473, 55–60. [Google Scholar] [CrossRef] [PubMed]
  3. Suga, M.; Akita, F.; Hirata, K.; Ueno, G.; Murakami, H.; Nakajima, Y.; Shimizu, T.; Yamashita, K.; Yamamoto, M.; Ago, H.; et al. Native structure of photosystem II at 1.95 Å resolution viewed by femtosecond X-ray pulses. Nature 2015, 517, 99–103. [Google Scholar] [CrossRef] [PubMed]
  4. Simons, K.; Ikonen, E. Functional rafts in cell membranes. Nature 1997, 387, 569–572. [Google Scholar] [CrossRef] [PubMed]
  5. Fang, R.H.; Jiang, Y.; Fang, J.C.; Zhang, L. Cell membrane-derived nanomaterials for biomedical applications. Biomaterials 2017, 128, 69–83. [Google Scholar] [CrossRef] [PubMed]
  6. Fang, R.H.; Kroll, A.V.; Gao, W.; Zhang, L. Cell membrane coating nanotechnology. Adv. Mater. 2018, 30, 1706759. [Google Scholar] [CrossRef]
  7. Khosravi, N.; Pishavar, E.; Baradaran, B.; Oroojalian, F.; Mokhtarzadeh, A. Stem cell membrane, stem cell-derived exosomes and hybrid stem cell camouflaged nanoparticles: A promising biomimetic nanoplatforms for cancer theranostics. J. Control. Release 2022, 348, 706–722. [Google Scholar] [CrossRef]
  8. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  9. Guo, D.; Shibuya, R.; Akiba, C.; Saji, S.; Kondo, T.; Nakamura, J. Active sites of nitrogen-doped carbon materials for oxygen reduction reaction clarified using model catalysts. Science 2016, 351, 361–365. [Google Scholar] [CrossRef]
  10. Saidul Islam, M.S.; Shudo, Y.; Hayami, S. Energy conversion and storage in fuel cells and super-capacitors from chemical modifications of carbon allotropes: State-of-art and prospect. Bull. Chem. Soc. Jpn. 2022, 95, 1–25. [Google Scholar] [CrossRef]
  11. Klyndyuk, A.I.; Chizhova, E.A.; Kharytonau, D.S.; Medvedev, D.A. Layered oxygen-deficient double perovskites as promising cathode materials for solid oxide fuel cells. Materials 2022, 15, 141. [Google Scholar] [CrossRef] [PubMed]
  12. Maeda, K.; Takeiri, F.; Kobayashi, G.; Matsuishi, S.; Ogino, H.; Ida, S.; Mori, T.; Uchimoto, Y.; Tanabe, S.; Hasegawa, T.; et al. Recent progress on mixed-anion materials for energy applications. Bull. Chem. Soc. Jpn. 2022, 95, 26–37. [Google Scholar] [CrossRef]
  13. Jakhar, M.; Kumar, A.; Ahluwalia, P.K.; Tankeshwar, K.; Pandey, R. Engineering 2D materials for photocatalytic water-splitting from a theoretical perspective. Materials 2022, 15, 2221. [Google Scholar] [CrossRef] [PubMed]
  14. Kinoshita, T. Highly efficient wideband solar energy conversion employing singlet-triplet transitions. Bull. Chem. Soc. Jpn. 2022, 95, 341–352. [Google Scholar] [CrossRef]
  15. Pastuszak, J.; Węgierek, P. Photovoltaic cell generations and current research directions for their development. Materials 2022, 15, 5542. [Google Scholar] [CrossRef]
  16. Murugan, S.; Lee, E.-C. Recent Advances in the synthesis and application of vacancy-ordered halide double perovskite materials for solar cells: A promising alternative to lead-based perovskites. Materials 2023, 16, 5275. [Google Scholar] [CrossRef]
  17. Desoky, M.M.H.; Caldera, F.; Brunella, V.; Ferrero, R.; Hoti, G.; Trotta, F. Cyclodextrins for lithium batteries applications. Materials 2023, 16, 5540. [Google Scholar] [CrossRef]
  18. Honda, M.; Suzuki, N. Toxicities of polycyclic aromatic hydrocarbons for aquatic animals. Int. J. Environ. Res. Public Health 2020, 17, 1363. [Google Scholar] [CrossRef]
  19. Chapman, A.; Ertekin, E.; Kubota, M.; Nagao, A.; Bertsch, K.; Macadre, A.; Tsuchiyama, T.; Masamura, T.; Takaki, S.; Komoda, R.; et al. Achieving a carbon neutral future through advanced functional materials and technologies. Bull. Chem. Soc. Jpn. 2022, 95, 73–103. [Google Scholar] [CrossRef]
  20. Qaidi, S.; Najm, H.M.; Abed, S.M.; Özkılıç, Y.O.; Al Dughaishi, H.; Alosta, M.; Sabri, M.M.S.; Alkhatib, F.; Milad, A. Concrete containing waste glass as an environmentally friendly aggregate: A review on fresh and mechanical characteristics. Materials 2022, 15, 6222. [Google Scholar] [CrossRef]
  21. Yagyu, J.; Islam, M.S.; Yasutake, H.; Hirayama, H.; Zenno, H.; Sugimoto, A.; Takagi, S.; Sekine, Y.; Ohira, S.; Hayami, S. Insights and further understanding of radioactive cesium removal using zeolite, Prussian blue and graphene oxide as adsorbents. Bull. Chem. Soc. Jpn. 2022, 95, 862–870. [Google Scholar] [CrossRef]
  22. Chen, S.; Liu, J.; Zhang, Q.; Teng, F.; McLellan, B.C. A critical review on deployment planning and risk analysis of carbon capture, utilization, and storage (CCUS) toward carbon neutrality. Renew. Sustain. Energy Rev. 2022, 167, 112537. [Google Scholar] [CrossRef]
  23. Mamun, M.R.A.; Yusuf, M.A.; Bhuyan, M.M.; Bhuiyan, M.S.H.; Arafath, M.A.; Uddin, M.N.; Soeb, M.J.A.; Almahri, A.; Rahman, M.M.; Karim, M.R. Acidity controlled desulfurization of biogas by using iron (III) and ferrosoferric (II, III) oxide. Bull. Chem. Soc. Jpn. 2022, 95, 1234–1241. [Google Scholar] [CrossRef]
  24. Saleh, T.S.; Badawi, A.K.; Salama, R.S.; Mostafa, M.M.M. Design and development of novel composites containing nickel ferrites supported on activated carbon derived from agricultural wastes and its application in water remediation. Materials 2023, 16, 2170. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, L.; Chong, H.L.H.; Moh, P.Y.; Albaqami, M.D.; Tighezza, A.M.; Qin, C.; Ni, X.; Cao, J.; Xu, X.; Yamauchi, Y. β-FeOOH nanospindles as chloride-capturing electrodes for electrochemical faradic deionization of saline water. Bull. Chem. Soc. Jpn. 2023, 96, 306–309. [Google Scholar] [CrossRef]
  26. Simonenko, N.P.; Glukhova, O.E.; Plugin, I.A.; Kolosov, D.A.; Nagornov, I.A.; Simonenko, T.L.; Varezhnikov, A.S.; Simonenko, E.P.; Sysoev, V.V.; Kuznetsov, N.T. The Ti0.2V1.8C MXene ink-prepared chemiresistor: From theory to tests with humidity versus VOCs. Chemosensors 2023, 11, 7. [Google Scholar] [CrossRef]
  27. Chowdhury, R.; Al Biruni, M.T.; Afia, A.; Hasan, M.; Islam, M.R.; Ahmed, T. Medical waste incineration fly ash as a mineral filler in dense bituminous course in flexible pavements. Materials 2023, 16, 5612. [Google Scholar] [CrossRef]
  28. Cabral, H.; Miyata, K.; Osada, K.; Kataoka, K. Block copolymer micelles in nanomedicine applications. Chem. Rev. 2018, 118, 6844–6892. [Google Scholar] [CrossRef]
  29. Fujita, Y.; Niizeki, T.; Fukumitsu, N.; Ariga, K.; Yamauchi, Y.; Malgras, V.; Kaneti, Y.V.; Liu, C.-H.; Hatano, K.; Suematsu, H.; et al. Mechanisms responsible for adsorption of molybdate ions on alumina for the production of medical radioisotopes. Bull. Chem. Soc. Jpn. 2022, 95, 129–137. [Google Scholar] [CrossRef]
  30. Islam, F.; Shohag, S.; Uddin, M.J.; Islam, M.R.; Nafady, M.H.; Akter, A.; Mitra, S.; Roy, A.; Emran, T.B.; Cavalu, S. Exploring the journey of zinc oxide nanoparticles (ZnO-NPs) toward biomedical applications. Materials 2022, 15, 2160. [Google Scholar] [CrossRef]
  31. Pradipta, A.R.; Michiba, H.; Kubo, A.; Fujii, M.; Tanei, T.; Morimoto, K.; Shimazu, K.; Tanaka, K. The second-generation click-to-sense probe for intraoperative diagnosis of breast cancer tissues based on acrolein targeting. Bull. Chem. Soc. Jpn. 2022, 95, 421–426. [Google Scholar] [CrossRef]
  32. Bhattacharya, T.; Soares, G.A.B.E.; Chopra, H.; Rahman, M.M.; Hasan, Z.; Swain, S.S.; Cavalu, S. Applications of phyto-nanotechnology for the treatment of neurodegenerative disorders. Materials 2022, 15, 804. [Google Scholar] [CrossRef] [PubMed]
  33. Sahayasheela, V.J.; Yu, Z.; Hirose, Y.; Pandian, G.N.; Bando, T.; Sugiyama, H. Inhibition of GLI-mediated transcription by cyclic pyrrole-imidazole polyamide in cancer stem cells. Bull. Chem. Soc. Jpn. 2022, 95, 693–699. [Google Scholar] [CrossRef]
  34. Quader, S.; Kataoka, K.; Cabral, H. Nanomedicine for brain cancer. Adv. Drug Deliv. Rev. 2022, 182, 114115. [Google Scholar] [CrossRef] [PubMed]
  35. Komiyama, M. Molecular mechanisms of the medicines for COVID-19. Bull. Chem. Soc. Jpn. 2022, 95, 1308–1317. [Google Scholar] [CrossRef]
  36. Larasati, L.; Lestari, W.W.; Firdaus, M. Dual-action Pt(IV) prodrugs and targeted delivery in metal-organic frameworks: Overcoming cisplatin resistance and improving anticancer activity. Bull. Chem. Soc. Jpn. 2022, 95, 1561–1577. [Google Scholar] [CrossRef]
  37. Niwa, T.; Tahara, T.; Chase, C.E.; Fang, F.G.; Nakaoka, T.; Irie, S.; Hayashinaka, E.; Wada, Y.; Mukai, H.; Masutomi, K.; et al. Synthesis of 11C-radiolabeled eribulin as a companion diagnostics PET tracer for brain glioblastoma. Bull. Chem. Soc. Jpn. 2023, 96, 283–290. [Google Scholar] [CrossRef]
  38. Zhang, E.; Zhu, Q.; Huang, J.; Liu, J.; Tan, G.; Sun, C.; Li, T.; Liu, S.; Li, Y.; Wang, H.; et al. Visually resolving the direct Z-scheme heterojunction in CdS@ZnIn2S4 hollow cubes for photocatalytic evolution of H2 and H2O2 from pure water. Appl. Catal. B Environ. 2021, 293, 120213. [Google Scholar] [CrossRef]
  39. Mori, H.; Yamada, Y.; Minagawa, Y.; Hasegawa, N.; Nishihara, Y. Effects of acyloxy groups in anthrabisthiadiazole-based semiconducting polymers on electronic properties, thin-film structure, and solar cell performance. Bull. Chem. Soc. Jpn. 2022, 95, 942–952. [Google Scholar] [CrossRef]
  40. Qi, Y.; Zhang, J.; Kong, Y.; Zhao, Y.; Chen, S.; Li, D.; Liu, W.; Chen, Y.; Xie, T.; Cui, J.; et al. Unraveling of cocatalysts photodeposited selectively on facets of BiVO4 to boost solar water splitting. Nat. Commun. 2022, 13, 484. [Google Scholar] [CrossRef]
  41. Charles-Blin, Y.; Kondo, T.; Wu, Y.; Bandow, S.; Awaga, K. Salt-assisted pyrolysis of covalent organic framework for controlled active nitrogen functionalities for oxygen reduction reaction. Bull. Chem. Soc. Jpn. 2022, 95, 972–977. [Google Scholar] [CrossRef]
  42. Takeuchi, Y.; Matsuzawa, K.; Nagai, T.; Ikegami, K.; Kuroda, Y.; Monden, R.; Ishihara, A. Fe, N-Doped SrTiO3 synthesized using pyrazine carboxylic acid-metal complexes: Application as an oxygen reduction catalyst for polymer electrolyte fuel cell cathodes in acidic media. Bull. Chem. Soc. Jpn. 2023, 96, 175–177. [Google Scholar] [CrossRef]
  43. Chen, K.; Xiao, J.; Vequizo, J.J.M.; Hisatomi, T.; Ma, Y.; Nakabayashi, M.; Takata, T.; Yamakata, A.; Shibata, N.; Domen, K. Overall water splitting by a SrTaO2N-based photocatalyst decorated with an Ir-promoted Ru-based cocatalyst. J. Am. Chem. Soc. 2023, 145, 3839–3843. [Google Scholar] [CrossRef] [PubMed]
  44. Yano, J.; Suzuk, K.; Hashimoto, C.; Tsutsumi, C.; Hayase, N.; Kitani, A. Trial fabrication of NADH-dependent enzymatic ethanol biofuel cell providing H2 gas as well as Electricity. Bull. Chem. Soc. Jpn. 2023, 96, 331–338. [Google Scholar] [CrossRef]
  45. Huang, G.; Hu, M.; Xu, X.; Alothman, A.A.; Mushab, M.S.S.; Ma, S.; Shen, P.K.; Zhu, J.; Yamauchi, Y. Optimizing heterointerface of Co2P–CoxOy Nanoparticles within a porous carbon network for deciphering superior water splitting. Small Struct. 2023, 4, 2200235. [Google Scholar] [CrossRef]
  46. Kumar, R.; Joanni, E.; Sahoo, S.; Shim, J.-J.; Tan, W.K.; Matsuda, A.; Singh, R.K. An overview of recent progress in nanostructured carbon-based supercapacitor electrodes: From zero to bi-dimensional materials. Carbon 2022, 193, 298–338. [Google Scholar] [CrossRef]
  47. Kumar, R.; Youssry, S.M.; Joanni, E.; Sahoo, S.; Kawamura, G.; Matsuda, A. Microwave-assisted synthesis of iron oxide homogeneously dispersed on reduced graphene oxide for high-performance supercapacitor electrodes. J. Energy Storage 2022, 56, 105896. [Google Scholar] [CrossRef]
  48. Das, H.T.; Dutta, S.; Balaji, T.E.; Das, N.; Das, P.; Dheer, N.; Kanojia, R.; Ahuja, P.; Ujjain, S.K. Recent rrends in carbon nanotube electrodes for flexible supercapacitors: A review of smart energy storage device assembly and performance. Chemosensors 2022, 10, 223. [Google Scholar] [CrossRef]
  49. Zhang, G.; Qiuhong Bai, Q.; Wang, X.; Li, C.; Uyama, H.; Shen, Y. Preparation and mechanism investigation of walnut shell-based hierarchical porous carbon for supercapacitors. Bull. Chem. Soc. Jpn. 2023, 96, 190–197. [Google Scholar] [CrossRef]
  50. Arumugam, B.; Mayakrishnan, G.; Subburayan Manickavasagam, S.K.; Kim, S.C.; Vanaraj, R. An overview of active electrode materials for the efficient high-performance supercapacitor application. Crystals 2023, 13, 1118. [Google Scholar] [CrossRef]
  51. Lakshmi, K.C.S.; Vedhanarayanan, B. High-performance supercapacitors: A comprehensive review on paradigm shift of conventional energy storage devices. Batteries 2023, 9, 202. [Google Scholar] [CrossRef]
  52. Gnawali, C.L.; Manandha, S.; Shahi, S.; Shrestha, R.G.; Adhikari, M.P.; Rajbhandari, R.; Pokharel, B.P.; Ma, R.; Ariga, K.; Shrestha, L.K. Nanoporous carbon materials from terminalia bellirica seed for iodine and methylene blue adsorption and high-performance supercapacitor applications. Bull. Chem. Soc. Jpn. 2023, 96, 572–581. [Google Scholar] [CrossRef]
  53. Sasai, R.; Fujimura, T.; Sato, H.; Nii, E.; Sugata, M.; Nakayashiki, Y.; Hoashi, H.; Moriyoshi, C.; Oishi, E.; Fujii, Y.; et al. Origin of selective nitrate removal by Ni2+–Al3+ layered double hydroxides in aqueous media and its application potential in seawater purification. Bull. Chem. Soc. Jpn. 2022, 95, 802–812. [Google Scholar] [CrossRef]
  54. Li, S.-L.; Chang, G.; Huang, C.Y.; Kinooka, K.; Chen, K.Y.; Fu, W.; Gong, G.; Yoshioka, T.; McKeown, N.B.; Hu, Y. 2,2′-Biphenol-based ultrathin microporous nanofilms for highly efficient molecular sieving separation. Angew. Chem. Int. Ed. 2022, 61, e202212816. [Google Scholar] [CrossRef] [PubMed]
  55. Ren, F.; He, R.; Ren, J.; Tao, F.; Yang, H.; Lv, H.; Ju, X. A friendly UV-responsive fluorine-free superhydrophobic coating for oil-water separation and dye degradation. Bull. Chem. Soc. Jpn. 2022, 95, 1091–1099. [Google Scholar] [CrossRef]
  56. Kushida, W.; Gonzales, R.R.; Shintani, T.; Matsuoka, A.; Nakagawa, K.; Yoshioka, T.; Hideto Matsuyama, H. Organic solvent mixture separation using fluorine-incorporated thin film composite reverse osmosis membrane. J. Mater. Chem. A 2022, 10, 4146–4156. [Google Scholar] [CrossRef]
  57. Wang, X.; Liu, Z.; Lu, J.; Teng, H.; Fukuda, H.; Qin, W.; Wei, T.; Liu, Y. Highly selective membrane for efficient separation of environmental determinands: Enhanced molecular imprinting in polydopamine-embedded porous sleeve. Chem. Eng. J. 2022, 449, 137825. [Google Scholar] [CrossRef]
  58. Kukobat, R.; Sakai, M.; Tanaka, H.; Otsuka, H.; Vallejos-Burgos, F.; Lastoskie, C.; Matsukata, M.; Sasaki, Y.; Yoshida, K.; Hayashi, T.; et al. Ultrapermeable 2D-channeled graphene-wrapped zeolite molecular sieving membranes for hydrogen separation. Sci. Adv. 2022, 8, eabl3521. [Google Scholar] [CrossRef]
  59. Zhang, T.; Cain, A.K.; Semenec, L.; Liu, L.; Hosokawa, Y.; Inglis, D.W.; Yalikun, Y.; Li, M. Microfluidic separation and enrichment of escherichia coli by size using viscoelastic flows. Anal. Chem. 2023, 95, 2561–2569. [Google Scholar] [CrossRef]
  60. Hung, H.-L.; Iizuka, T.; Deng, X.; Lyu, Q.; Hsu, C.-H.; Oe, N.; Lin, L.-C.; Hosono, N.; Kang, D.-Y. Engineering gas separation property of metal–organic framework membranes via polymer insertion. Sep. Purif. Technol. 2023, 310, 123115. [Google Scholar] [CrossRef]
  61. Salahuddin, B.; Masud, M.K.; Aziz, S.; Liu, C.-H.; Amiralian, N.; Ashok, A.; Hossain, M.A.; Park, H.; Wahab, M.A.; Amin, M.A.; et al. p-Carrageenan gel modified mesoporous gold chronocoulometric sensor for ultrasensitive detection of microRNA. Bull. Chem. Soc. Jpn. 2022, 95, 198–207. [Google Scholar] [CrossRef]
  62. Dong, L.; Wang, M.; Wu, J.; Zhu, C.; Shi, J.; Morikawa, H. Stretchable, adhesive, self-healable, and conductive hydrogel-based deformable triboelectric nanogenerator for energy harvesting and human motion sensing. ACS Appl. Mater. Interfaces 2022, 14, 9126–9137. [Google Scholar] [CrossRef] [PubMed]
  63. Han, X.; Wang, S.; Liu, M.; Liu, L. A cucurbit[6]uril-based supramolecular assembly as a multifunctional material for the detection and removal of organic explosives and antibiotics. Bull. Chem. Soc. Jpn. 2022, 95, 1445–1452. [Google Scholar] [CrossRef]
  64. Masuda, Y. Recent advances in SnO2 nanostructure based gas sensors. Sens. Actuat. B Chem. 2022, 364, 131876. [Google Scholar] [CrossRef]
  65. Saito, S.; Haraga, T.; Marumo, K.; Sato, Y.; Nakano, Y.; Tasaki-Handa, Y.; Shibukawa, M. Americium(III)/curium(III) complete separation and sensitive fluorescence detection by capillary and gel electrophoresis using emissive hexadentate/octadentate polyaminocarboxylate ligands. Bull. Chem. Soc. Jpn. 2023, 96, 223–225. [Google Scholar] [CrossRef]
  66. Kalyana Sundaram, S.d.; Hossain, M.M.; Rezki, M.; Ariga, K.; Tsujimura, S. Enzyme cascade electrode reactions with nanomaterials and their applicability towards biosensor and biofuel cells. Biosensors 2023, 13, 1018. [Google Scholar] [CrossRef]
  67. Anggraini, L.E.; Rahmawati, H.; Nasution, M.A.F.; Jiwanti, P.K.; Einaga, Y.; Ivandini, T.A. Development of an acrylamide biosensor using guanine and adenine as biomarkers at boron-doped diamond electrodes: Integrated molecular docking and experimental studies. Bull. Chem. Soc. Jpn. 2023, 96, 420–428. [Google Scholar] [CrossRef]
  68. Nakagawa, Y.; Tsujimura, S.; Zelsmann, M.; Zebda, A. Hierarchical structure of gold and carbon electrode for Bilirubin oxidase-biocathode. Biosensors 2023, 13, 482. [Google Scholar] [CrossRef]
  69. Maeki, M.; Uno, S.; Niwa, A.; Okada, Y.; Tokeshi, M. Microfluidic technologies and devices for lipid nanoparticle-based RNA delivery. J. Control. Release 2022, 344, 80–96. [Google Scholar] [CrossRef]
  70. Tiburcius, S.; Krishnan, K.; Patel, V.; Netherton, J.; Sathish, C.I.; Weidenhofer, J.; Yang, J.-H.; Verrills, N.M.; Karakoti, A.; Vinu, A. Triple surfactant assisted synthesis of novel core-shell mesoporous silica nanoparticles with high surface area for drug delivery for prostate cancer. Bull. Chem. Soc. Jpn. 2022, 95, 331–340. [Google Scholar] [CrossRef]
  71. Yokoo, H.; Oba, M.; Uchida, S. Cell-penetrating peptides: Emerging tools for mRNA delivery. Pharmaceutics 2022, 14, 78. [Google Scholar] [CrossRef] [PubMed]
  72. Su, C.-H.; Soendoro, A.; Okayama, S.; Rahmania, F.J.; Nagai, T.; Imae, T.; Tsutsumiuchi, K.; Kawai, N. Drug release stimulated by magnetic field and light on magnetite- and carbon dot-loaded carbon nanohorn. Bull. Chem. Soc. Jpn. 2022, 95, 582–594. [Google Scholar] [CrossRef]
  73. Zhao, Y.; Das, S.; Sekine, T.; Mabuchi, H.; Irie, T.; Sakai, J.; Wen, D.; Zhu, W.; Ben, T.; Negishi, Y. Record ultralarge-pores, low density three-dimensional covalent organic Framework for controlled drug delivery. Angew. Chem. Int. Ed. 2023, 62, e202300172. [Google Scholar] [CrossRef] [PubMed]
  74. Younis, M.A.; Sato, Y.; Elewa, Y.H.A.; Kon, Y.; Harashima, H. Self-homing nanocarriers for mRNA delivery to the activated hepatic stellate cells in liver fibrosis. J. Control. Release 2023, 353, 685–698. [Google Scholar] [CrossRef] [PubMed]
  75. Rajan, R.; Kumar, N.; Zhao, D.; Dai, X.; Kawamoto, K.; Matsumura, K. Polyampholyte-based polymer hydrogels for the long-term storage, protection and delivery of therapeutic proteins. Adv. Healthc. Mater. 2023, 12, 2203253. [Google Scholar] [CrossRef] [PubMed]
  76. Khan, M.S.; Baskoy, S.A.; Yang, C.; Hong, J.; Chae, J.; Ha, H.; Lee, S.; Tanaka, M.; Choi, Y.; Choi, J. Lipid-based colloidal nanoparticles for applications in targeted vaccine delivery. Nanoscale Adv. 2023, 5, 1853–1869. [Google Scholar] [CrossRef] [PubMed]
  77. Ito, H.; Segawa, Y.; Murakami, K.; Itami, K. Polycyclic arene synthesis by annulative π-extension. J. Am. Chem. Soc. 2019, 141, 3–10. [Google Scholar] [CrossRef]
  78. Fan, W.; Matsuno, T.; Han, Y.; Wang, X.; Zhou, Q.; Isobe, H.; Wu, J. Synthesis and chiral resolution of twisted carbon nanobelts. J. Am. Chem. Soc. 2021, 143, 15924–15929. [Google Scholar] [CrossRef]
  79. Tanaka, T. Synthesis of novel heteronanographenes via fold-in approach. Bull. Chem. Soc. Jpn. 2022, 95, 602–610. [Google Scholar] [CrossRef]
  80. Miera, G.G.; Matsubara, S.; Kono, H.; Murakami, K.; Itami, K. Synthesis of octagon-containing molecular nanocarbons. Chem. Sci. 2022, 13, 1848–1868. [Google Scholar] [CrossRef]
  81. Mori, K. C(sp3)–H Bond functionalization mediated by hydride a shift/cyclization system. Bull. Chem. Soc. Jpn. 2022, 95, 296–305. [Google Scholar] [CrossRef]
  82. Sugiyama, M.; Akiyama, M.; Yonezawa, Y.; Komaguchi, K.; Higashi, M.; Nozaki, K.; Okazoe, T. Electron in a cube: Synthesis and characterization of perfluorocubane as an electron acceptor. Science 2022, 377, 756–759. [Google Scholar] [CrossRef] [PubMed]
  83. Segawa, Y. Nonplanar aromatic hydrocarbons: Design and synthesis of highly strained structures. Bull. Chem. Soc. Jpn. 2022, 95, 1600–1610. [Google Scholar] [CrossRef]
  84. Hirano, K. Copper-catalyzed electrophilic amination: An umpolung strategy for new C–N bond formations. Bull. Chem. Soc. Jpn. 2023, 96, 198–207. [Google Scholar] [CrossRef]
  85. Toya, M.; Omine, T.; Ishiwari, F.; Saeki, A.; Ito, H.; Itami, K. Expanded [2,1][n]carbohelicenes with 15- and 17-benzene rings. J. Am. Chem. Soc. 2023, 145, 11553–11565. [Google Scholar] [CrossRef] [PubMed]
  86. Tsubaki, N.; Wang, Y.; Yang, G.; He, Y. Rational design of novel reaction pathways and tailor-made catalysts for value-added chemicals synthesis from CO2 hydrogenation. Bull. Chem. Soc. Jpn. 2023, 96, 291–302. [Google Scholar] [CrossRef]
  87. Hossain, S.; Niihori, Y.; Nair, L.V.; Kumar, B.; Kurashige, W.; Negishi, Y. Alloy clusters: Precise synthesis and mixing effects. Acc. Chem. Res. 2018, 51, 3114–3124. [Google Scholar] [CrossRef]
  88. Tokoro, H.; Nakabayashi, K.; Nagashima, S.; Song, Q.; Yoshikiyo, M.; Ohkoshi, S. Optical properties of epsilon iron oxide nanoparticles in the millimeter- and terahertz-wave regions. Bull. Chem. Soc. Jpn. 2022, 95, 538–552. [Google Scholar] [CrossRef]
  89. Taniguchi, T.; Nurdiwijayanto, L.; Ma, R.; Sasaki, T. Chemically exfoliated inorganic nanosheets for nanoelectronics. Appl. Phys. Rev. 2022, 9, 021313. [Google Scholar] [CrossRef]
  90. Jiříčková, A.; Jankovský, O.; Sofer, Z.; Sedmidubský, D. Synthesis and applications of graphene oxide. Materials 2022, 15, 920. [Google Scholar] [CrossRef]
  91. Yadav, N.; Gaikwad, R.P.; Mishra, V.; Gawande, M.B. Synthesis and photocatalytic Applications of functionalized carbon quantum dots. Bull. Chem. Soc. Jpn. 2022, 95, 1638–1679. [Google Scholar] [CrossRef]
  92. Wu, D.; Kusada, K.; Yamamoto, T.; Toriyama, T.; Matsumura, S.; Kawaguchi, S.; Kubota, Y.; Kitagawa, H. Platinum-group-metal high-entropy-alloy nanoparticles. J. Am. Chem. Soc. 2020, 142, 13833–13838. [Google Scholar] [CrossRef] [PubMed]
  93. Peera, S.G.; Koutavarapu, R.; Chao, L.; Singh, L.; Murugadoss, G.; Rajeshkhanna, G. 2D MXene nanomaterials as electrocatalysts for hydrogen evolution reaction (HER): A review. Micromachines 2022, 13, 1499. [Google Scholar] [CrossRef] [PubMed]
  94. Naganthran, A.; Verasoundarapandian, G.; Khalid, F.E.; Masarudin, M.J.; Zulkharnain, A.; Nawawi, N.M.; Karim, M.; Che Abdullah, C.A.; Ahmad, S.A. Synthesis, characterization and biomedical application of silver nanoparticles. Materials 2022, 15, 427. [Google Scholar] [CrossRef]
  95. Minamihara, H.; Kusada, K.; Yamamoto, T.; Toriyama, T.; Murakami, Y.; Matsumura, S.; Kumara, L.S.R.; Sakata, O.; Kawaguchi, S.; Kubota, Y.; et al. Continuous-flow chemical synthesis for sub-2 nm ultra-multielement alloy nanoparticles consisting of group IV to XV elements. J. Am. Chem. Soc. 2023, 145, 17136–17142. [Google Scholar] [CrossRef]
  96. Okamoto, K.; Imoto, H.; Naka, K. Silsesquioxane cage-fused siloxane rings as a novel class of inorganic-based host molecules. Bull. Chem. Soc. Jpn. 2023, 96, 84–89. [Google Scholar] [CrossRef]
  97. Kitagawa, S.; Kitaura, R.; Noro, S. Functional porous coordination polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef]
  98. Shan, Y.; Zhang, G.; Yin, W.; Pang, H.; Xu, Q. Recent progress in Prussian blue/Prussian blue analogue-derived metallic compounds. Bull. Chem. Soc. Jpn. 2022, 95, 230–260. [Google Scholar] [CrossRef]
  99. Bieniek, A.; Terzyk, A.P.; Wiśniewski, M.; Roszek, K.; Kowalczyk, P.; Sarkisov, L.; Keskin, S.; Kaneko, K. MOF materials as therapeutic agents, drug carriers, imaging agents and biosensors in cancer biomedicine: Recent advances and perspectives. Prog. Mater. Sci. 2021, 117, 100743. [Google Scholar] [CrossRef]
  100. Yam, V.W.-W.; Cheng, Y.-H. Stimuli-responsive and switchable platinum(II) complexes and their applications in memory storage. Bull. Chem. Soc. Jpn. 2022, 95, 846–854. [Google Scholar] [CrossRef]
  101. Kim, M.; Xin, R.; Earnshaw, J.; Tang, J.; Hill, J.P.; Ashok, A.; Nanjundan, A.K.; Kim, J.; Young, C.; Sugahara, Y.; et al. MOF-derived nanoporous carbons with diverse tunable nanoarchitectures. Nat. Protoc. 2022, 17, 2990–3027. [Google Scholar] [CrossRef] [PubMed]
  102. Sánchez-González, E.; Tsang, M.Y.; Troyano, J.; Craig, G.A.; Furukawa, S. Assembling metal–organic cages as porous materials. Chem. Soc. Rev. 2022, 51, 4876–4889. [Google Scholar] [CrossRef] [PubMed]
  103. Sarango-Ramírez, M.K.; Donoshita, M.; Yoshida, Y.; Lim, D.-W.; Kitagawa, H. Cooperative proton and Li-ion conduction in a 2D-layered MOF via mechanical insertion of lithium halides. Angew. Chem. Int. Ed. 2023, 62, e202301284. [Google Scholar] [CrossRef] [PubMed]
  104. Ohki, Y. Biomimetic transition metal cluster complexes: Synthetic studies and applications in the reduction of inert small molecules. Bull. Chem. Soc. Jpn. 2023, 96, 639–648. [Google Scholar] [CrossRef]
  105. Hayashi, R.; Tashiro, S.; Asakura, M.; Mitsui, S.; Shionoya, M. Effector-dependent structural transformation of a crystalline framework with allosteric effects on molecular recognition ability. Nat. Commun. 2023, 14, 4490. [Google Scholar] [CrossRef] [PubMed]
  106. Fujiwara, A.; Wang, J.; Hiraide, S.; Götz, A.; Miyahara, M.T.; Hartmann, M.; Apeleo Zubiri, B.; Spiecker, E.; Vogel, N.; Watanabe, S. Fast gas-adsorption kinetics in supraparticle-based MOF packings with hierarchical porosity. Adv. Mater. 2023, 35, 2305980. [Google Scholar] [CrossRef] [PubMed]
  107. Datta, S.; Kato, Y.; Higashiharaguchi, S.; Aratsu, K.; Isobe, A.; Saito, T.; Prabhu, D.D.; Kitamoto, Y.; Hollamby, M.J.; Smith, A.J.; et al. Self-assembled poly-catenanes from supramolecular toroidal building blocks. Nature 2020, 583, 400–405. [Google Scholar] [CrossRef]
  108. Takezawa, H.; Shitozawa, K.; Fujita, M. Enhanced reactivity of twisted amides inside a molecular cage. Nat. Chem. 2020, 12, 574–578. [Google Scholar] [CrossRef]
  109. Sasaki, Y.; Kubota, R.; Minami, T. Molecular self-assembled chemosensors and their arrays. Coord. Chem. Rev. 2021, 429, 213607. [Google Scholar] [CrossRef]
  110. Hamada, K.; Shimoyama, D.; Hirao, T.; Haino, T. Chiral supramolecular polymer formed via host-guest complexation of an octaphosphonate biscavitand and a chiral diammonium guest. Bull. Chem. Soc. Jpn. 2022, 95, 621–627. [Google Scholar] [CrossRef]
  111. Zhang, X.-J.; Morishita, D.; Aoki, T.; Itoh, Y.; Yano, K.; Araoka, F.; Aida, T. Anomalous chiral transfer: Supramolecular polymerization in a chiral medium of a mesogenic molecule. Chem. Asian J. 2022, 17, e202200223. [Google Scholar] [CrossRef] [PubMed]
  112. Bezrukov, A.; Galyametdinov, Y. Activation and switching of supramolecular chemical signals in multi-output microfluidic devices. Micromachines 2022, 13, 1778. [Google Scholar] [CrossRef] [PubMed]
  113. Matsuno, T.; Isobe, H. Trapped yet free inside the tube: Supramolecular chemistry of molecular peapods. Bull. Chem. Soc. Jpn. 2023, 96, 406–419. [Google Scholar] [CrossRef]
  114. Kubota, R. Supramolecular–polymer composite hydrogels: From In Situ network observation to functional properties. Bull. Chem. Soc. Jpn. 2023, 96, 802–812. [Google Scholar] [CrossRef]
  115. Otsuka, C.; Takahashi, S.; Isobe, A.; Saito, T.; Aizawa, T.; Tsuchida, R.; Yamashita, S.; Harano, K.; Hanayama, H.; Shimizu, N.; et al. Supramolecular polymer polymorphism: Spontaneous helix–helicoid transition through dislocation of hydrogen-bonded p-rosettes. J. Am. Chem. Soc. 2023, 145, 22563–22576. [Google Scholar] [CrossRef]
  116. Ogoshi, T. Supramolecular assemblies and polymer recognition based on polygonal and pillar-shaped macrocycles “pillar[n]arenes”. Polym. J. 2023, 55, 1247–1260. [Google Scholar] [CrossRef]
  117. Kamigaito, M.; Ando, T.; Sawamoto, M. Metal-catalyzed living radical polymerization. Chem. Rev. 2001, 101, 3689–3746. [Google Scholar] [CrossRef]
  118. Kitao, T.; Zhang, Y.; Kitagawa, S.; Wan, B.; Uemura, T. Hybridization of MOFs and polymers. Chem. Soc. Rev. 2017, 46, 3108–3133. [Google Scholar] [CrossRef]
  119. Nandihalli, N.; Liu, C.J.; Mori, T. Polymer based thermoelectric nanocomposite materials and devices: Fabrication and characteristics. Nano Energy 2020, 78, 105186. [Google Scholar] [CrossRef]
  120. Zhang, D.; Liu, D.; Ubukata, T.; Seki, T. Unconventional approaches to light-promoted dynamic surface morphing on polymer films. Bull. Chem. Soc. Jpn. 2022, 95, 138–162. [Google Scholar] [CrossRef]
  121. Zhang, W.; You, L.; Meng, X.; Wang, B.; Lin, D. Recent advances on conducting polymers based nanogenerators for energy harvesting. Micromachines 2021, 12, 1308. [Google Scholar] [CrossRef] [PubMed]
  122. Tsuchii, Y.; Menda, T.; Hwang, S.; Yasuda, T. Photovoltaic properties of p-conjugated polymers based on fused cyclic Imide and amide skeletons. Bull. Chem. Soc. Jpn. 2023, 96, 90–94. [Google Scholar] [CrossRef]
  123. Hatakeyama-Sato, K.; Oyaizu, K. Redox: Organic robust radicals and their polymers for energy conversion/storage devices. Chem. Rev. 2023, 123, 11336–11391. [Google Scholar] [CrossRef] [PubMed]
  124. Hosokawa, S.; Nagao, A.; Hashimoto, Y.; Matsune, A.; Okazoe, T.; Suzuki, C.; Wada, H.; Kakiuchi, T.; Tsuda, A. Non-Isocyanate polyurethane synthesis by polycondensation of alkylene and arylene bis(fluoroalkyl) bis(carbonate)s with diamines. Bull. Chem. Soc. Jpn. 2023, 96, 663–670. [Google Scholar] [CrossRef]
  125. Iizuka, T.; Sano, H.; Le Ouay, B.; Hosono, N.; Uemura, T. An approach to MOFaxanes by threading ultralong polymers through metal–organic framework microcrystals. Nat. Commun. 2023, 14, 3241. [Google Scholar] [CrossRef]
  126. Kawaguchi, D. Aggregation states, thermal molecular motion and carrier properties in functional polymer thin films. Polym. J. 2023, 55, 1237–1245. [Google Scholar] [CrossRef]
  127. Akiba, U.; Minaki, D.; Anzai, J. Host-guest chemistry in layer-by-layer assemblies containing calix[n]arenes and cucurbit[n]urils: A Review. Polymers 2018, 10, 130. [Google Scholar] [CrossRef]
  128. Minamiki, T.; Ichikawa, Y.; Kurita, R. The power of assemblies at interfaces: Nanosensor platforms based on synthetic receptor membranes. Sensors 2020, 20, 2228. [Google Scholar] [CrossRef]
  129. Yang, X.; Biswas, S.K.; Han, J.; Tanpichai, S.; Li, M.-C.; Chen, C.; Zhu, S.; Das, A.K.; Yano, H. Surface and interface engineering for nanocellulosic Advanced materials. Adv. Mater. 2021, 33, 2002264. [Google Scholar] [CrossRef]
  130. Zhu, H. Interfacial preparation of ferroelectric polymer nanostructures for electronic applications. Polym. J. 2021, 53, 877–886. [Google Scholar] [CrossRef]
  131. Miyabe, K.; Aoki, K. Moment analysis of solute permeation kinetics at an interface of mixed micelles of anionic and nonionic surfactants. Bull. Chem. Soc. Jpn. 2022, 95, 1715–1722. [Google Scholar] [CrossRef]
  132. Zhang, Z.; Zeng, J.; Groll, J.; Matsusaki, M. Layer-by-layer assembly methods and their biomedical applications. Biomater. Sci. 2022, 10, 4077–4094. [Google Scholar] [CrossRef] [PubMed]
  133. Makiura, R. Creation of metal–organic framework nanosheets by the Langmuir-Blodgett technique. Coord. Chem. Rev. 2022, 469, 214650. [Google Scholar] [CrossRef]
  134. Kawasaki, Y.; Nakagawa, M.; Ito, T.; Imura, Y.; Wang, K.-H.; Kawai, T. Chiral transcription from chiral Au nanowires to self-assembled monolayers of achiral azobenzene derivatives. Bull. Chem. Soc. Jpn. 2022, 95, 1006–1010. [Google Scholar] [CrossRef]
  135. Yao, M.S.; Otake, K.; Kitagawa, S. Interface chemistry of conductive crystalline porous thin films. Trends Chem. 2023, 5, 588–604. [Google Scholar] [CrossRef]
  136. Akamatsu, M.; Sakai, K.; Sakai, H. Dynamic control of interfacial properties and self-assembly with photoirradiation. Chem. Lett. 2023, 52, 573–581. [Google Scholar] [CrossRef]
  137. Mitsudo, K.; Kurimoto, Y.; Yoshioka, K.; Suga, S. Miniaturization and combinatorial approach in organic electrochemistry. Chem. Rev. 2018, 118, 5985–5999. [Google Scholar] [CrossRef]
  138. Yamamoto, K.; Imaoka, T.; Tanabe, M.; Kambe, T. New horizon of nanoparticle and cluster catalysis with dendrimers. Chem. Rev. 2020, 120, 1397–1437. [Google Scholar] [CrossRef]
  139. Pan, Z.-Z.; Lv, W.; Yang, Q.-H.; Nishihara, H. Aligned macroporous monoliths by ice-templating. Bull. Chem. Soc. Jpn. 2022, 95, 611–620. [Google Scholar] [CrossRef]
  140. Razaq, A.; Bibi, F.; Zheng, X.; Papadakis, R.; Jafri, S.H.M.; Li, H. Review on graphene-, graphene oxide-, reduced graphene oxide-based flexible composites: From fabrication to applications. Materials 2022, 15, 1012. [Google Scholar] [CrossRef]
  141. Adschiri, T.; Takami, S.; Umetsu, M.; Ohara, S.; Naka, T.; Minami, K.; Hojo, D.; Togashi, T.; Arita, T.; Taguchi, M.; et al. Supercritical hydrothermal reactions for material synthesis. Bull. Chem. Soc. Jpn. 2023, 96, 133–147. [Google Scholar] [CrossRef]
  142. Zeybek, Ö.; Özkılıç, Y.O.; Karalar, M.; Çelik, A.İ.; Qaidi, S.; Ahmad, J.; Burduhos-Nergis, D.D.; Burduhos-Nergis, D.P. Influence of replacing cement with waste glass on mechanical properties of concrete. Materials 2022, 15, 7513. [Google Scholar] [CrossRef] [PubMed]
  143. Yang, J.; Yamato, M.; Shimizu, T.; Sekine, H.; Ohashi, K.; Kanzaki, M.; Ohki, T.; Nishida, K.; Okano, T. Reconstruction of functional tissues with cell sheet engineering. Biomaterials 2007, 28, 5033–5043. [Google Scholar] [CrossRef] [PubMed]
  144. Mijanović, O.; Pylaev, T.; Nikitkina, A.; Artyukhova, M.; Branković, A.; Peshkova, M.; Bikmulina, P.; Turk, B.; Bolevich, S.; Avetisov, S.; et al. Tissue engineering meets nanotechnology: Molecular mechanism modulations in cornea regeneration. Micromachines 2021, 12, 1336. [Google Scholar] [CrossRef] [PubMed]
  145. Tamura, T.; Inoue, M.; Yoshimitsu, Y.; Hashimoto, I.; Ohashi, N.; Tsumura, K.; Suzuki, K.; Watanabe, T.; Hohsaka, T. Chemical synthesis and cell-free expression of thiazoline ring-bridged cyclic Peptides and their properties on biomembrane permeability. Bull. Chem. Soc. Jpn. 2022, 95, 359–366. [Google Scholar] [CrossRef]
  146. Rabiee, N.; Ahmadi, S.; Akhavan, O.; Luque, R. Silver and gold nanoparticles for antimicrobial purposes against multi-drug resistance bacteria. Materials 2022, 15, 1799. [Google Scholar] [CrossRef] [PubMed]
  147. Yoo, D.; Nakamura, M.; Kanjo, M.; Gon, M.; Watanabe, H.; Kita, H.; Tanaka, K. Facile preparation of near infrared-luminescent protein complexes with conjugated polymers consisting of boron azobenzene units. Bull. Chem. Soc. Jpn. 2023, 96, 659–662. [Google Scholar] [CrossRef]
  148. Zhou, P.; Xing, R.; Li, Q.; Li, J.; Yuan, C.; Yan, X. Steering phase-separated droplets to control fibrillar network evolution of supramolecular peptide hydrogels. Matter 2023, 6, 1945–1963. [Google Scholar] [CrossRef]
  149. Miyagawa, A.; Hagiya, K.; Nagatomo, S.; Nakatani, K. Protein adsorption on carboxy-functionalized microparticles revealed by zeta potential and absorption spectroscopy measurements. Bull. Chem. Soc. Jpn. 2023, 96, 759–765. [Google Scholar] [CrossRef]
  150. Li, Z.; Yu, F.; Xu, X.; Wang, T.; Fei, J.; Hao, J.; Li, J. Photozyme-catalyzed ATP generation based on ATP synthase-reconstituted nanoarchitectonics. J. Am. Chem. Soc. 2023, 145, 20907–20912. [Google Scholar] [CrossRef]
  151. Hieda, M.; Tsujimura, K.; Kinoshita, M.; Matsumori, N. Formation of a tight complex between amphidinol and sterols in lipid bilayers revealed by short-range energy transfer. Bull. Chem. Soc. Jpn. 2022, 95, 1753–1759. [Google Scholar] [CrossRef]
  152. Inamasu, R.; Yamaguchi, H.; Arai, T.; Chang, J.; Kuramochi, M.; Mio, K.; Sasaki, Y.V. Observation of molecular motions in polymer thin films by laboratory grazing incidence diffracted X-ray blinking. Polym. J. 2023, 55, 703–709. [Google Scholar] [CrossRef]
  153. Liu, C.; Morimoto, N.; Jiang, L.; Kawahara, S.; Noritomi, T.; Yokoyama, H.; Mayumi, K.; Ito, K. Tough hydrogels with rapid self-reinforcement. Science 2021, 372, 1078–1081. [Google Scholar] [CrossRef] [PubMed]
  154. Shichijo, K.; Watanabe, M.; Hisaeda, Y.; Shimakoshi, H. Development of visible light-driven hybrid catalysts composed of earth abundant metal ion modified TiO2 and B12 complex. Bull. Chem. Soc. Jpn. 2022, 95, 1016–1024. [Google Scholar] [CrossRef]
  155. Uchida, J.; Soberats, B.; Gupta, M.; Kato, T. Advanced functional liquid crystals. Adv. Mater. 2022, 34, 2109063. [Google Scholar] [CrossRef] [PubMed]
  156. Roopsung, N.; An, T.L.H.; Sugawara, A.; Asoh, T.; Hsu, Y.-I.; Uyama, H. Switchable stiffness of composite hydrogels triggered by thermoresponsive phase-change particles. Bull. Chem. Soc. Jpn. 2023, 96, 636–638. [Google Scholar] [CrossRef]
  157. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W.; Fujishima, A.; Terashima, C. Enhanced photocatalytic degradation activity using the V2O5/RGO Composite. Nanomaterials 2023, 13, 338. [Google Scholar] [CrossRef]
  158. Li, Z.; Nakamura, K.; Kobayashi, N. Solvent-free mechanochemical synthesis of strongly luminescent Eu(III) hybrid materials using tetramethylammonium salts. Bull. Chem. Soc. Jpn. 2023, 96, 816–823. [Google Scholar] [CrossRef]
  159. Liu, M.; Ishida, Y.; Ebina, Y.; Sasaki, T.; Hikima, T.; Takata, M.; Aida, T. An anisotropic hydrogel with electrostatic repulsion between cofacially aligned nanosheets. Nature 2015, 517, 68–72. [Google Scholar] [CrossRef]
  160. Kasuya, N.; Tsurumi, J.; Okamoto, T.; Watanabe, S.; Takeya, J. Two-dimensional hole gas in organic semiconductors. Nat. Mater. 2021, 20, 1401–1406. [Google Scholar] [CrossRef]
  161. Nishijima, A.; Kametani, Y.; Uemura, T. Reciprocal regulation between MOFs and polymers. Coord. Chem. Rev. 2022, 466, 214601. [Google Scholar] [CrossRef]
  162. Jung, T.A.; Schlittler, R.R.; Gimzewski, J.K.; Tang, H.; Joachim, C. Controlled room-temperature positioning of individual molecules: Molecular flexure and motion. Science 1996, 271, 181–184. [Google Scholar] [CrossRef]
  163. Sugimoto, Y.; Pou, P.; Abe, M.; Jelinek, P.; Pérez, R.; Morita, S.; Custance, Ó. Chemical identification of individual surface atoms by atomic force microscopy. Nature 2007, 446, 64–67. [Google Scholar] [CrossRef] [PubMed]
  164. Soe, W.-H.; Srivastava, S.; Joachim, C. Train of single molecule-gears. J. Phys. Chem. Lett. 2019, 10, 6462–6467. [Google Scholar] [CrossRef] [PubMed]
  165. Kawai, S.; Krejčí, O.; Nishiuchi, T.; Sahara, K.; Kodama, T.; Pawlak, R.; Meyer, E.; Kubo, T.; Foster, A.S. Three-dimensional graphene nanoribbons as a framework for molecular assembly and local probe chemistry. Sci. Adv. 2020, 6, eaay8913. [Google Scholar] [CrossRef] [PubMed]
  166. Xing, J.; Takeuchi, K.; Kamei, K.; Nakamuro, T.; Harano, K.; Nakamura, E. Atomic-number (Z)-correlated atomic sizes for deciphering electron microscopic molecular images. Proc. Natl. Acad. Sci. USA 2022, 119, e2114432119. [Google Scholar] [CrossRef] [PubMed]
  167. Tada, K.; Hinuma, Y.; Ichikawa, S.; Tanaka, S. Investigation of the interaction between Au and brookite TiO2 using transmission electron microscopy and density functional theory. Bull. Chem. Soc. Jpn. 2023, 96, 373–380. [Google Scholar] [CrossRef]
  168. Kimura, K.; Miwa, K.; Imada, H.; Imai-Imada, M.; Kawahara, S.; Takeya, J.; Kawai, M.; Galperin, M.; Kim, Y. Selective triplet exciton formation in a single molecule. Nature 2019, 570, 210–213. [Google Scholar] [CrossRef]
  169. Kawai, S.; Sang, H.; Kantorovich, L.; Takahashi, K.; Nozaki, K.; Ito, S. An endergonic synthesis of single Sondheimer–Wong diyne by local probe chemistry. Angew. Chem. Int. Ed. 2020, 59, 10842. [Google Scholar] [CrossRef]
  170. Imahori, H. Molecular photoinduced charge separation: Fundamentals and application. Bull. Chem. Soc. Jpn. 2023, 96, 339–352. [Google Scholar] [CrossRef]
  171. Ito, S.; Akama, H.; Kojima, K.M.; McKenzie, I.; Kuwahata, K.; Tachikawa, M. Muon spin rotation (μSR) for characterizing radical addition to C=S in xanthene-9-thione and thioxanthene-9-thione. Bull. Chem. Soc. Jpn. 2023, 96, 461–464. [Google Scholar] [CrossRef]
  172. Liu, X.; Farahi, G.; Chiu, C.-L.; Papic, Z.; Watanabe, K.; Taniguchi, T.; Zaletel, M.P.; Yazdani, A. Visualizing broken symmetry and topological defects in a quantum Hall ferromagnet. Science 2022, 375, 321–326. [Google Scholar] [CrossRef] [PubMed]
  173. Yamaguchi, T.; Higa, S.; Yoshida, K.; Sumitani, K.; Kurisaki, T. Structure of aqueous scandium(III) nitrate solution by large-angle X-ray scattering combined with empirical potential refinement modeling, X-ray absorption fine structure, and discrete variational Xα calculations. Bull. Chem. Soc. Jpn. 2022, 95, 673–679. [Google Scholar] [CrossRef]
  174. Itoh, T.; Procházka, M.; Dong, Z.-C.; Ji, W.; Yamamoto, Y.S.; Zhang, Y.; Ozaki, Y. Toward a new era of SERS and TERS at the nanometer scale: From fundamentals to innovative applications. Chem. Rev. 2023, 123, 1552–1634. [Google Scholar] [CrossRef] [PubMed]
  175. Kameda, Y.; Arai, N.; Amo, Y.; Usuki, T.; Han, J.; Watanabe, H.; Umebayashi, Y.; Tsuzuki, S.; Ikeda, K.; Otomo, T. Neutron diffraction with 34S/natS isotopic substitution method on the solvation structure of S8 molecule in concentrated CS2 solutions. Bull. Chem. Soc. Jpn. 2022, 95, 1481–1485. [Google Scholar] [CrossRef]
  176. Lostao, A.; Lim, K.S.; Pallarés, M.C.; Ptak, A.; Marcuello, C. Recent advances in sensing the inter-biomolecular interactions at the nanoscale—A comprehensive review of AFM-based force spectroscopy. Int. J. Biol. Macromol. 2023, 238, 124089. [Google Scholar] [CrossRef]
  177. Yamashita, Y.; Tsurumi, J.; Ohno, M.; Fujimoto, R.; Kumagai, S.; Kurosawa, T.; Okamoto, T.; Takeya, J.; Watanabe, S. Efficient molecular doping of polymeric semiconductors driven by anion exchange. Nature 2019, 572, 634–638. [Google Scholar] [CrossRef]
  178. Kobayashi, H.; Takeuchi, K.; Morinaga, Y.; Honda, H.; Yamamoto, M.; Odanaka, Y.; Inagaki, M. Inter-spin interactions of 1D chains of different-sized nitroxide radicals incorporated in the organic 1D nanochannels of tris(o-phenylenedioxy)cyclotriphosphazene. Bull. Chem. Soc. Jpn. 2022, 95, 981–988. [Google Scholar] [CrossRef]
  179. Tsuji, H. Carbon-bridged oligo(phenylenevinylene)s that reveal cryogenic phenomena at room temperature. Bull. Chem. Soc. Jpn. 2022, 95, 657–662. [Google Scholar] [CrossRef]
  180. Fan, L.-B.; Shu, C.-C.; Dong, D.; He, J.; Henriksen, N.E.; Nori, F. Quantum coherent control of a single molecular-polariton rotation. Phys. Rev. Lett. 2023, 130, 043604. [Google Scholar] [CrossRef]
  181. Wooten, B.L.; Iguchi, R.; Tang, P.; Kang, J.S.; Uchida, K.; Bauer, G.E.W.; Heremans, J.P. Electric field–dependent phonon spectrum and heat conduction in ferroelectrics. Sci. Adv. 2021, 9, eadd7194. [Google Scholar] [CrossRef] [PubMed]
  182. Ariga, K.; Li, M.; Richards, G.J.; Hill, J.P. Nanoarchitectonics: A conceptual paradigm for design and synthesis of dimension-conrolled functional nanomaterials. J. Nanosci. Nanotechnol. 2011, 11, 1–13. [Google Scholar] [CrossRef] [PubMed]
  183. Ariga, K.; Ji, Q.; Nakanishi, W.; Hill, J.P.; Aono, M. Nanoarchitectonics: A new materials horizon for nanotechnolog. Mater. Horiz. 2015, 2, 406–413. [Google Scholar] [CrossRef]
  184. Cao, L.; Huang, Y.; Parakhonskiy, B.; Skirtach, A.G. Nanoarchitectonics beyond perfect order—Not quite perfect but quite useful. Nanoscale 2022, 14, 15964–16002. [Google Scholar] [CrossRef] [PubMed]
  185. Gupta, D.; Varghese, B.S.; Suresh, M.; Panwar, C.; Gupta, T.K. Nanoarchitectonics: Functional nanomaterials and nanostructures—A review. J. Nanopart. Res. 2022, 24, 196. [Google Scholar] [CrossRef]
  186. Feynman, R.P. There’s plenty of room at the bottom. Calif. Inst. Technol. J. Eng. Sci. 1960, 4, 23–36. [Google Scholar]
  187. Roukes, M. Plenty of room, indeed. Sci. Am. 2001, 285, 48–57. [Google Scholar] [CrossRef]
  188. Ariga, K.; Ji, Q.; Hill, J.; Bando, Y.; Aono, M. Forming nanomaterials as layered functional structures toward materials nanoarchitectonics. NPG Asia Mater. 2012, 4, e17. [Google Scholar] [CrossRef]
  189. Ariga, K.; Minami, K.; Ebara, M.; Nakanishi, J. What are the emerging concepts and challenges in NANO? Nanoarchitectonics, hand-operating nanotechnology and mechanobiology. Polym. J. 2016, 48, 371–389. [Google Scholar] [CrossRef]
  190. Ariga, K. Nanoarchitectonics: What’s coming next after nanotechnology? Nanoscale Horiz. 2021, 6, 364–378. [Google Scholar] [CrossRef]
  191. Ariga, K. Molecular nanoarchitectonics: Unification of nanotechnology and molecular/materials science. Beilstein J. Nanotechnol. 2023, 14, 434–453. [Google Scholar] [CrossRef] [PubMed]
  192. Ariga, K.; Li, J.; Fei, J.; Ji, Q.; Hill, J.P. Nanoarchitectonics for dynamic functional materials from atomic-/molecular-level manipulation to macroscopic action. Adv. Mater. 2016, 28, 1251–1286. [Google Scholar] [CrossRef] [PubMed]
  193. Ariga, K.; Jia, X.; Song, J.; Hill, J.P.; Leong, D.T.; Jia, Y.; Li, J. Nanoarchitectonics beyond self-assembly: Challenges to create bio-like hierarchic organization. Angew. Chem. Int. Ed. 2020, 59, 15424–15446. [Google Scholar] [CrossRef] [PubMed]
  194. Ariga, K.; Nishikawa, M.; Mori, T.; Takeya, J.; Shrestha, L.K.; Hill, J.P. Self-assembly as a key player for materials nanoarchitectonics. Sci. Technol. Adv. Mater. 2019, 20, 51–95. [Google Scholar] [CrossRef] [PubMed]
  195. Li, T.; Lu, X.-M.; Zhang, M.-R.; Hu, K.; Li, Z. Peptide-based nanomaterials: Self-assembly, properties and applications. Bioact. Mater. 2022, 11, 268–282. [Google Scholar] [CrossRef] [PubMed]
  196. Mu, Q.; Liu, R.; Kimura, H.; Li, J.; Jiang, H.; Zhang, X.; Yu, Z.; Sun, X.; Algadi, H.; Guo, Z.; et al. Supramolecular self-assembly synthesis of hemoglobin-like amorphous CoP@N, P-doped carbon composites enable ultralong stable cycling under high-current density for lithium-ion battery anodes. Adv. Compos. Hybrid Mater. 2023, 6, 23. [Google Scholar] [CrossRef]
  197. Ariga, K. Nanoarchitectonics: A navigator from materials to life. Mater. Chem. Front. 2017, 1, 208–211. [Google Scholar] [CrossRef]
  198. Aono, M.; Ariga, K. The way to nanoarchitectonics and the way of nanoarchitectonics. Adv. Mater. 2016, 28, 989–992. [Google Scholar] [CrossRef]
  199. Ariga, K. Liquid interfacial nanoarchitectonics: Molecular machines, organic semiconductors, nanocarbons, stem cells, and others. Curr. Opin. Colloid Interface Sci. 2023, 63, 101656. [Google Scholar] [CrossRef]
  200. Shen, X.; Song, J.; Kawakami, K.; Ariga, K. Molecule-to-material-to-bio nanoarchitectonics with biomedical fullerene nanoparticles. Materials 2022, 15, 5404. [Google Scholar] [CrossRef]
  201. Ariga, K.; Fakhrullin, R. Materials nanoarchitectonics from atom to living cell: A method for everything. Bull. Chem. Soc. Jpn. 2022, 95, 774–795. [Google Scholar] [CrossRef]
  202. Ariga, K. Nanoarchitectonics: Method for everything in material science. Bull. Chem. Soc. Jpn. 2024; in press. [Google Scholar] [CrossRef]
  203. Laughlin, R.B.; Pines, D. The theory of everything. Proc. Natl. Acad. Sci. USA 2000, 97, 28–31. [Google Scholar] [CrossRef] [PubMed]
  204. Khan, A.H.; Ghosh, S.; Pradhan, B.; Dalui, A.; Shrestha, L.K.; Acharya, S.; Ariga, K. Two-dimensional (2D) nanomaterials towards electrochemical nanoarchitectonics in energy-related applications. Bull. Chem. Soc. Jpn. 2017, 90, 627–648. [Google Scholar] [CrossRef]
  205. Oaki, Y.; Sato, K. Nanoarchitectonics for conductive polymers using solid and vapor phases. Nanoscale Adv. 2022, 4, 2773–2781. [Google Scholar] [CrossRef] [PubMed]
  206. Nguyen, N.T.; Lebastard, C.; Wilmet, M.; Dumait, N.; Renaud, A.; Cordier, S.; Ohashi, N.; Uchikoshi, T.; Grasset, F. A review on functional nanoarchitectonics nanocomposites based on octahedral metal atom clusters (Nb6, Mo6, Ta6, W6, Re6): Inorganic 0D and 2D powders and films. Sci. Technol. Adv. Mater. 2022, 23, 547–578. [Google Scholar] [CrossRef] [PubMed]
  207. Datta, K.K.R. Exploring the self-cleaning facets of fluorinated graphene nanoarchitectonics: Progress and perspectives. ChemNanoMat 2023, 9, e202300135. [Google Scholar] [CrossRef]
  208. Zhang, Z.-P.; Xia, H. Nanoarchitectonics and applications of gallium-based liquid metal micro- and nanoparticles. ChemNanoMat 2023, 9, e202300078. [Google Scholar] [CrossRef]
  209. Hikichi, R.; Tokura, Y.; Igarashi, Y.; Imai, H.; Oaki, Y. Fluorine-free substrate-independent superhydrophobic Coatings by nanoarchitectonics of polydispersed 2D materials. Bull. Chem. Soc. Jpn. 2023, 96, 766–774. [Google Scholar] [CrossRef]
  210. Trifoi, A.R.; Matei, E.; Râpă, M.; Berbecaru, A.-C.; Panaitescu, C.; Banu, I.; Doukeh, R. Coprecipitation nanoarchitectonics for the synthesis of magnetite: A review of mechanism and characterization. React. Kinet. Mech. Catal. 2023, 136, 2835–2874. [Google Scholar] [CrossRef]
  211. Guan, X.; Li, Z.; Geng, X.; Lei, Z.; Karakoti, A.; Wu, T.; Kumar, P.; Yi, J.; Vinu, A. Emerging trends of carbon-based quantum dots: Nanoarchitectonics and applications. Small 2023, 19, 2207181. [Google Scholar] [CrossRef] [PubMed]
  212. Jiang, H.; Gao, J.; Zhang, X.; Guo, N. Composite micro-nanoarchitectonics of MMT-SiO2: Space charge characteristics under tensile state. Polymers 2021, 13, 4354. [Google Scholar] [CrossRef]
  213. Xing, Z.; Zhang, C.; Xue, N.; Li, Z.; Li, F.; Wan, X.; Guo, S.; Hao, J. High-frequency surface insulation strength with nanoarchitectonics of disiloxane modified polyimide films. Polymers 2022, 14, 146. [Google Scholar] [CrossRef] [PubMed]
  214. Jia, J.; Wu, D.; Ren, Y.; Lin, J. Nanoarchitectonics of illite-based materials: Effect of metal oxides lntercalation on the mechanical properties. Nanomaterials 2022, 12, 997. [Google Scholar] [CrossRef] [PubMed]
  215. Chen, Q.; Li, H. Nanoarchitectonics of carbon nanostructures: Sodium dodecyl sulfonate @ sodium chloride system. Nanomaterials 2022, 12, 1652. [Google Scholar] [CrossRef] [PubMed]
  216. Hakim, M.L.; Hanif, A.; Alam, T.; Islam, M.T.; Arshad, H.; Soliman, M.S.; Albadran, S.M.; Islam, M.S. Ultrawideband polarization-independent nanoarchitectonics: A prfect metamaterial absorber for visible and infrared optical window applications. Nanomaterials 2022, 12, 2849. [Google Scholar] [CrossRef] [PubMed]
  217. Lin, Y.-F.; Lai, Y.-R.; Sung, H.-L.; Chung, T.-W.; Lin, K.-Y.A. Design of amine-modified Zr–Mg mixed oxide aerogel nanoarchitectonics with dual lewis acidic and basic sites for CO2/propylene oxide cycloaddition reactions. Nanomaterials 2022, 12, 3442. [Google Scholar] [CrossRef]
  218. Ling, L.; Wu, C.; Xing, F.; Memon, S.A.; Sun, H. Recycling nanoarchitectonics of graphene oxide from carbon fiber Reinforced polymer by the electrochemical method. Nanomaterials 2022, 12, 3657. [Google Scholar] [CrossRef]
  219. Qiu, Z.; Jinschek, J.R.; Gouma, P.-I. Two-step solvothermal process for nanoarchitectonics of metastable hexagonal WO3 nanoplates. Crystals 2023, 13, 690. [Google Scholar] [CrossRef]
  220. Kim, D.; Gu, M.; Park, M.; Kim, T.; Kim, B.-S. Layer-by-layer assembly for photoelectrochemical nanoarchitectonics. Mol. Syst. Des. Eng. 2019, 4, 65–77. [Google Scholar] [CrossRef]
  221. Zhang, L.; Wang, T.; Shen, Z.; Liu, M. Chiral nanoarchitectonics: Towards the design, self-assembly, and function of nanoscale chiral twists and helices. Adv. Mater. 2016, 28, 1044–1059. [Google Scholar] [CrossRef] [PubMed]
  222. Ariga, K.; Shionoya, M. Nanoarchitectonics for coordination asymmetry and related chemistry. Bull. Chem. Soc. Jpn. 2021, 94, 839–859. [Google Scholar] [CrossRef]
  223. Marin, E.; Tapeinos, C.; Sarasua, J.R.; Larrañaga, A. Exploiting the layer-by-layer nanoarchitectonics for the fabrication of polymer capsules: A toolbox to provide multifunctional properties to target complex pathologies. Adv. Colloid Interface Sci. 2022, 304, 102680. [Google Scholar] [CrossRef] [PubMed]
  224. Li, Z.; Zhao, C.; Lin, X.; Ouyang, G.; Liu, M. Stepwise solution-interfacial nanoarchitectonics for assembled film with full-color and white-light circularly polarized luminescence. ACS Appl. Mater. Interfaces 2023, 15, 31077–31086. [Google Scholar] [CrossRef] [PubMed]
  225. Parbat, D.; Jana, N.; Dhar, M.; Mann, U. Reactive multilayer coating as versatile nanoarchitectonics for customizing various bioinspired liquid wettabilities. ACS Appl. Mater. Interfaces 2022, 15, 25232–25247. [Google Scholar] [CrossRef]
  226. Zhang, X.; Yang, P. Role of graphitic carbon in g-C3N4 nanoarchitectonics towards efficient photocatalytic reaction kinetics: A review. Carbon 2023, 216, 118584. [Google Scholar] [CrossRef]
  227. Haketa, Y.; Yamasumi, K.; Maeda, H. p-Electronic ion pairs: Building blocks for supramolecular nanoarchitectonics via ip-ip interactions. Chem. Soc. Rev. 2023, 52, 7170–7196. [Google Scholar] [CrossRef]
  228. Huang, S.-Y.; Hsieh, P.-Y.; Chung, C.-J.; Chou, C.-M.; He, J.-L. Nanoarchitectonics for ultrathin gold films deposited on collagen fabric by high-power impulse magnetron sputtering. Nanomaterials 2022, 12, 1627. [Google Scholar] [CrossRef]
  229. Béres, K.A.; Homonnay, Z.; Bereczki, L.; Dürvanger, Z.; Petruševski, V.M.; Farkas, A.; Kótai, L. Crystal nanoarchitectonics and characterization of the octahedral iron(III)–nitrate complexes with isomer dimethylurea ligands. Crystals 2023, 13, 1019. [Google Scholar] [CrossRef]
  230. Lahmidi, S.; Anouar, E.H.; Ettahiri, W.; El Hafi, M.; Lazrak, F.; Alanazi, M.M.; Alanazi, A.S.; Hefnawy, M.; Essassi, E.M.; Mague, J.T. Nanoarchitectonics and molecular docking of 4-(dimethylamino)pyridin-1-ium 2-3 methyl-4-oxo-pyri-do[1,2-a]pyrimidine-3-carboxylate. Crystals 2023, 13, 1333. [Google Scholar] [CrossRef]
  231. Hecht, S. Welding, organizing, and planting organic molecules on substrate surfaces—Promising approaches towards nanoarchitectonics from the bottom up. Angew. Chem. Int. Ed. 2003, 42, 24–26. [Google Scholar] [CrossRef] [PubMed]
  232. Ramanathan, M.; Shrestha, L.K.; Mori, T.; Ji, Q.; Hill, J.P.; Ariga, K. Amphiphile nanoarchitectonics: From basic physical chemistry to advanced applications. Phys. Chem. Chem. Phys. 2013, 15, 10580–10611. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, Y.; Niu, D.; Ouyang, G.; Liu, M. Double helical p-aggregate nanoarchitectonics for amplified circularly polarized luminescence. Nat. Commun. 2022, 13, 1710. [Google Scholar] [CrossRef] [PubMed]
  234. Akamatsu, M. Inner and interfacial environmental nanoarchitectonics of supramolecular assemblies formed by amphiphiles: From emergence to application. J. Oleo Sci. 2023, 72, 105–116. [Google Scholar] [CrossRef] [PubMed]
  235. Pahal, S.; Boranna, R.; Tripathy, A.; Goudar, V.S.; Veetil, V.T.; Kurapati, R.; Prashanth, G.R.; Vemula, P.K. Nanoarchitectonics for free-standing polyelectrolyte multilayers films: Exploring the flipped surfaces. ChemNanoMat 2023, 9, e202200462. [Google Scholar] [CrossRef]
  236. Huang, L.; Yang, J.; Asakura, Y.; Shuai, Q.; Yamauchi, Y. Nanoarchitectonics of hollow covalent organic frameworks: Synthesis and applications. ACS Nano 2023, 17, 8918–8934. [Google Scholar] [CrossRef] [PubMed]
  237. Deepak, D.; Soin, N.; Roy, S.S. Optimizing the efficiency of triboelectric nanogenerators by surface nanoarchitectonics of graphene-based electrodes: A review. Mater. Today Commun. 2023, 34, 105412. [Google Scholar] [CrossRef]
  238. Musa, A.; Hakim, M.L.; Alam, T.; Islam, M.T.; Alshammari, A.S.; Mat, K.; Almalki, S.H.; Islam, M.S. Polarization independent metamaterial absorber with anti-reflection coating nanoarchitectonics for visible and infrared window applications. Materials 2022, 15, 3733. [Google Scholar] [CrossRef]
  239. Conti Nibali, V.; D’Angelo, G.; Arena, A.; Ciofi, C.; Scandurra, G.; Branca, C. TiO2 Nanoparticles dispersion in block-copolymer aqueous solutions: Nanoarchitectonics for self-assembly and aggregation. J. Funct. Biomater. 2022, 13, 39. [Google Scholar] [CrossRef]
  240. Zeng, L.; Liu, Z.; Huang, J.; Wang, X.; Guo, H.; Li, W.-H. Anti-fouling performance of hydrophobic hydrogels with unique surface hydrophobicity and nanoarchitectonics. Gels 2022, 8, 407. [Google Scholar] [CrossRef]
  241. Nayak, A.; Unayama, S.; Tai, S.; Tsuruoka, T.; Waser, R.; Aono, M.; Valov, I.; Hasegawa, T. Nanoarchitectonics for controlling the number of dopant atoms in solid electrolyte nanodots. Adv. Mater. 2018, 30, 1703261. [Google Scholar] [CrossRef] [PubMed]
  242. Eguchi, M.; Nugraha, A.S.; Rowan, A.E.; Shapter, J.; Yamauchi, Y. Adsorchromism: Molecular nanoarchitectonics at 2D nanosheets—Old chemistry for advanced chromism. Adv. Sci. 2021, 8, 2100539. [Google Scholar] [CrossRef] [PubMed]
  243. Yao, B.; Sun, H.; He, Y.; Wang, S.; Liu, X. Recent advances in the photoreactions triggered by porphyrin-based triplet–triplet annihilation upconversion systems: Molecular innovations and nanoarchitectonics. Int. J. Mol. Sci. 2022, 23, 8041. [Google Scholar] [CrossRef] [PubMed]
  244. Li, M.; Wu, Z.; Tian, Y.; Pan, F.; Gould, T.; Zhang, S. Nanoarchitectonics of two-dimensional electrochromic materials: Achievements and future challenges. Adv. Mater. Technol. 2023, 8, 2200917. [Google Scholar] [CrossRef]
  245. Ogawa, S. Aqueous sugar-based amphiphile systems: Recent advances in phase behavior and nanoarchitectonics. J. Oleo Sci. 2023, 72, 489–499. [Google Scholar] [CrossRef] [PubMed]
  246. Qiu, Y.; Zhou, X.; Tang, X.; Hao, Q.; Chen, M. Micro spectrometers based on materials nanoarchitectonics. Materials 2023, 16, 2253. [Google Scholar] [CrossRef]
  247. Zhang, X.; Yang, P. CsPbX3 (X = Cl, Br, and I) Nanocrystals in substrates toward stable photoluminescence: Nanoarchitectonics, properties, and applications. Langmuir 2023, 39, 11188–11212. [Google Scholar] [CrossRef]
  248. Lun-Fu, A.V.; Bubenchikov, A.M.; Bubenchikov, M.A.; Ovchinnikov, V.A. Numerical simulation of interaction between Kr+ ion and rotating C60 fullerene towards for nanoarchitectonics of fullerene materials. Crystals 2021, 11, 1204. [Google Scholar] [CrossRef]
  249. Li, X.; Weng, L.; Wang, H.; Wang, X. Nanoarchitectonics of BN/AgNWs/epoxy composites with high thermal conductivity and electrical insulation. Polymers 2021, 13, 4417. [Google Scholar] [CrossRef]
  250. Yang, K.; Qin, G.; Wang, L.; Zhao, M.; Lu, C. Theoretical nanoarchitectonics of GaN nanowires for ultraviolet irradiation-dependent electromechanical properties. Materials 2023, 16, 1080. [Google Scholar] [CrossRef]
  251. Tang, R.; Li, G.; Hu, X.; Gao, N.; Li, J.; Huang, K.; Kang, J.; Zhang, R. Micro-nanoarchitectonics of Ga2O3/GaN core-shell rod arrays for high-performance broadband ultraviolet photodetection. Crystals 2023, 13, 366. [Google Scholar] [CrossRef]
  252. Gao, K.; Wu, N.; Ji, B.; Liu, J. A film electrode upon nanoarchitectonics of bacterial cellulose and conductive fabric for forehead electroencephalogram measurement. Sensors 2023, 23, 7887. [Google Scholar] [CrossRef] [PubMed]
  253. Howorka, S. DNA Nanoarchitectonics: Assembled DNA at interfaces. Langmuir 2013, 29, 7344–7353. [Google Scholar] [CrossRef] [PubMed]
  254. Pandeeswar, M.; Senanayak, S.P.; Govindaraju, T. Nanoarchitectonics of small molecule and DNA for ultrasensitive detection of mercury. ACS Appl. Mater. Interfaces 2016, 8, 30362–30371. [Google Scholar] [CrossRef] [PubMed]
  255. Zou, Q.; Liu, K.; Abbas, M.; Yan, X. Peptide-modulated self-assembly of chromophores toward biomimetic light-harvesting nanoarchitectonics. Adv. Mater. 2016, 28, 1031–1043. [Google Scholar] [CrossRef] [PubMed]
  256. Stulz, E. Nanoarchitectonics with porphyrin functionalized DNA. Acc. Chem. Res. 2017, 50, 823–831. [Google Scholar] [CrossRef] [PubMed]
  257. Komiyama, M.; Yoshimoto, K.; Sisido, M.; Ariga, K. Chemistry can make strict and fuzzy controls for bio-systems: DNA Nanoarchitectonics and cell-macromolecular nanoarchitectonics. Bull. Chem. Soc. Jpn. 2017, 90, 967–1004. [Google Scholar] [CrossRef]
  258. Jia, Y.; Yan, X.; Li, J. Schiff base mediated dipeptide assembly toward nanoarchitectonics. Angew. Chem. Int. Ed. 2022, 61, e202207752. [Google Scholar] [CrossRef]
  259. Jia, X.; Chen, J.; Lv, W.; Li, H.; Ariga, K. Engineering dynamic and interactive biomaterials using material nanoarchitectonics for modulation of cellular behaviors. Cell Rep. Phys. Sci. 2023, 4, 101251. [Google Scholar] [CrossRef]
  260. Chang, R.; Zhao, L.; Xing, R.; Li, J.; Yan, X. Functional chromopeptide nanoarchitectonics: Molecular design, self-assembly and biological applications. Chem. Soc. Rev. 2023, 52, 2688–2712. [Google Scholar] [CrossRef]
  261. Nakanishi, W.; Minami, K.; Shrestha, L.K.; Ji, Q.; Hill, J.P.; Ariga, K. Bioactive nanocarbon assemblies: Nanoarchitectonics and applications. Nano Today 2014, 9, 378–394. [Google Scholar] [CrossRef]
  262. Liu, Q.; Li, H.; Yu, B.; Meng, Z.; Zhang, X.; Li, J.; Zheng, L. DNA-based dissipative assembly toward nanoarchitectonics. Adv. Funct. Mater. 2022, 32, 2201196. [Google Scholar] [CrossRef]
  263. Ferhan, A.R.; Park, S.; Park, H.; Tae, H.; Jackman, J.A.; Cho, N.-J. Lipid nanoparticle technologies for nucleic acid delivery: A nanoarchitectonics perspective. Adv. Funct. Mater. 2022, 32, 2203669. [Google Scholar] [CrossRef]
  264. Komiyama, M. Cyclodextrins as eminent constituents in nanoarchitectonics for drug delivery systems. Beilstein J. Nanotechnol. 2023, 14, 218–232. [Google Scholar] [CrossRef] [PubMed]
  265. Shen, X.; Song, J.; Sevencan, C.; Leong, D.T.; Ariga, K. Bio-interactive nanoarchitectonics with two-dimensional materials and environments. Sci. Technol. Adv. Mater. 2022, 23, 199–224. [Google Scholar] [CrossRef] [PubMed]
  266. Hu, W.; Shi, J.; Lv, W.; Jia, X.; Ariga, K. Regulation of stem cell fate and function by using bioactive materials with nanoarchitectonics for regenerative medicine. Sci. Technol. Adv. Mater. 2022, 23, 393–412. [Google Scholar] [CrossRef]
  267. Wu, M.; Liu, J.; Wang, X.; Zeng, H. Recent advances in antimicrobial surfaces via tunable molecular interactions: Nanoarchitectonics and bioengineering applications. Curr. Opin. Colloid Interface Sci. 2023, 66, 101707. [Google Scholar] [CrossRef]
  268. Piorecka, K.; Kurjata, J.; Stanczyk, W.A. Nanoarchitectonics: Complexes and conjugates of platinum drugs with silicon containing nanocarriers. An overview. Int. J. Mol. Sci. 2021, 22, 9264. [Google Scholar] [CrossRef]
  269. Zhang, R.; Wang, Y.; Yang, G. DNA–lysozyme nanoarchitectonics: Quantitative investigation on charge inversion and compaction. Polymers 2022, 14, 1377. [Google Scholar] [CrossRef]
  270. Czarnecka, E.; Nowaczyk, J.; Prochoń, M.; Masek, A. Nanoarchitectonics for biodegradable superabsorbent based on carboxymethyl starch and chitosan cross-linked with vanillin. Int. J. Mol. Sci. 2022, 23, 5386. [Google Scholar] [CrossRef]
  271. Kalinova, R.; Mladenova, K.; Petrova, S.; Doumanov, J.; Dimitrov, I. Nanoarchitectonics of spherical nucleic acids with biodegradable polymer cores: Synthesis and evaluation. Materials 2022, 15, 8917. [Google Scholar] [CrossRef] [PubMed]
  272. Osetrov, K.; Uspenskaya, M.; Zaripova, F.; Olekhnovich, R. Nanoarchitectonics of a skin-adhesive hydrogel based on the gelatin resuscitation fluid Gelatinol®. Gels 2023, 9, 330. [Google Scholar] [CrossRef] [PubMed]
  273. Wang, C.; Wang, H.; Na, J.; Yao, Y.; Azhar, A.; Yan, X.; Qi, J.; Yamauchi, Y.; Li, J. 0D–1D hybrid nanoarchitectonics: Tailored design of FeCo@N–C yolk–shell nanoreactors with dual sites for excellent Fenton-like catalysis. Chem. Sci. 2021, 12, 15418–15422. [Google Scholar] [CrossRef] [PubMed]
  274. Huang, C.; Qin, P.; Luo, Y.; Ruan, Q.; Liu, L.; Wu, Y.; Li, Q.; Xu, Y.; Liu, R.; Chu, P.K. Recent progress and perspective of cobalt-based catalysts for water splitting: Design and nanoarchitectonics. Mater. Today Energy 2022, 23, 100911. [Google Scholar] [CrossRef]
  275. Arulraj, A.; Murugesan, P.K.; Rajkumar, C.; Zamorano, A.T.; Mangalaraja, R.V. Nanoarchitectonics of layered metal chalcogenides-based ternary electrocatalyst for water splitting. Energies 2023, 16, 1669. [Google Scholar] [CrossRef]
  276. Kumar, A.; Choudhary, P.; Chhabra, T.; Kaur, H.; Kumar, A.; Qamar, M.; Krishnan, V. Frontier nanoarchitectonics of graphitic carbon nitride based plasmonic photocatalysts and photoelectrocatalysts for energy, environment and organic reactions. Mater. Chem. Front. 2023, 7, 1197–1247. [Google Scholar] [CrossRef]
  277. Jiang, B.; Guo, Y.; Sun, F.; Wang, S.; Kang, Y.; Xu, X.; Zhao, J.; You, J.; Eguchi, M.; Yamauch, Y.; et al. Nanoarchitectonics of metallene materials for electrocatalysis. ACS Nano 2023, 17, 13017–13043. [Google Scholar] [CrossRef]
  278. Lee, G.; Hossain, M.A.; Yoon, M.; Jhung, S.H. Nanoarchitectonics of metal–organic frameworks having hydroxy group for adsorption, catalysis, and sensing. J. Ind. Eng. Chem. 2023, 119, 181–192. [Google Scholar] [CrossRef]
  279. Sharma, D.; Choudhary, P.; Kumar, S.; Krishnan, V. Transition metal phosphide nanoarchitectonics for versatile organic catalysis. Small 2023, 19, 2207053. [Google Scholar] [CrossRef]
  280. Wang, Y.; Zhu, S.; He, S.; Lu, J.; Liu, J.; Lu, H.; Song, D.; Luo, Y. Nanoarchitectonics of Ni/CeO2 catalysts: The effect of pretreatment on the low-temperature steam reforming of glycerol. Nanomaterials 2022, 12, 816. [Google Scholar] [CrossRef]
  281. Cao, H.; Li, H.; Liu, L.; Xue, K.; Niu, X.; Hou, J.; Chen, L. Salt-templated nanoarchitectonics of CoSe2-NC nanosheets as an efficient bifunctional oxygen electrocatalyst for water splitting. Int. J. Mol. Sci. 2022, 23, 5239. [Google Scholar] [CrossRef] [PubMed]
  282. Liao, S.; Lin, L.; Huang, J.; Jing, X.; Chen, S.; Li, Q. Microorganism-templated nanoarchitectonics of hollow TiO2-SiO2 microspheres with enhanced photocatalytic activity for degradation of methyl orange. Nanomaterials 2022, 12, 1606. [Google Scholar] [CrossRef] [PubMed]
  283. Ishihara, S.; Labuta, J.; Van Rossom, W.; Ishikawa, D.; Minami, K.; Hil, J.P.; Ariga, K. Porphyrin-based sensor nanoarchitectonics in diverse physical detection modes. Phys. Chem. Chem. Phys. 2014, 16, 9713–9746. [Google Scholar] [CrossRef] [PubMed]
  284. Komiyama, M.; Mori, T.; Ariga, K. Molecular imprinting: Materials nanoarchitectonics with molecular information. Bull. Chem. Soc. Jpn. 2018, 91, 1075–1111. [Google Scholar] [CrossRef]
  285. Jadhav, R.W.; Khobrekar, P.P.; Bugde, S.T.; Bhosale, S.V. Nanoarchitectonics of neomycin-derived fluorescent carbon dots for selective detection of Fe3+ ions. Anal. Methods 2022, 14, 3289–3298. [Google Scholar] [CrossRef] [PubMed]
  286. Ma, K.; Yang, L.; Liu, J.; Liu, J. Electrochemical sensor nanoarchitectonics for sensitive detection of uric acid in human whole blood based on screen-printed carbon electrode equipped with vertically-ordered mesoporous silica-nanochannel film. Nanomaterials 2022, 12, 1157. [Google Scholar] [CrossRef] [PubMed]
  287. Nishat, Z.S.; Hossain, T.; Islam, M.N.; Phan, H.-P.; Wahab, M.A.; Moni, M.A.; Salomon, C.; Amin, M.A.; Sina, A.A.I.; Hossain, M.S.A.; et al. Hydrogel nanoarchitectonics: An evolving paradigm for ultrasensitive biosensing. Small 2022, 18, 2107571. [Google Scholar] [CrossRef]
  288. Vaghasiya, J.V.; Mayorga-Martinez, C.C.; Pumera, M. Wearable sensors for telehealth based on emerging materials and nanoarchitectonics. npj Flex. Electron. 2023, 7, 26. [Google Scholar] [CrossRef]
  289. Kim, S.K.; Lee, J.U.; Jeon, M.J.; Kim, S.-K.; Hwang, S.-H.; Honge, M.E.; Sim, S.J. Bio-conjugated nanoarchitectonics with dual-labeled nanoparticles for a colorimetric and fluorescent dual-mode serological lateral flow immunoassay sensor in detection of SARS-CoV-2 in clinical samples. RSC Adv. 2023, 13, 27225–27232. [Google Scholar] [CrossRef]
  290. Joshi, V.; Hussain, S.; Dua, S.; Arora, N.; Mir, S.H.; Rydzek, G.; Senthilkumar, T. Oligomer sensor nanoarchitectonics for “turn-on” fluorescence detection of cholesterol at the nanomolar level. Molecules 2022, 27, 2856. [Google Scholar] [CrossRef]
  291. Singh, V.; Thamizhanban, A.; Lalitha, K.; Subbiah, D.K.; Rachamalla, A.K.; Rebaka, V.P.; Banoo, T.; Kumar, Y.; Sridharan, V.; Ahmad, A.; et al. Self-assembling nanoarchitectonics of twisted nanofibers of fluorescent amphiphiles as chemo-resistive sensor for methanol detection. Gels 2023, 9, 442. [Google Scholar] [CrossRef] [PubMed]
  292. Ariga, K.; Ji, Q.; Mori, T.; Naito, M.; Yamauchi, Y.; Abe, H.; Hill, J.P. Enzyme nanoarchitectonics: Organization and device application. Chem. Soc. Rev. 2013, 42, 6322–6345. [Google Scholar] [CrossRef] [PubMed]
  293. Vuk, D.; Radovanović-Perić, F.; Mandić, V.; Lovrinčević, V.; Rath, T.; Panžić, I.; Le-Cunff, J. Synthesis and nanoarchitectonics of novel squaraine derivatives for organic photovoltaic devices. Nanomaterials 2022, 12, 1206. [Google Scholar] [CrossRef] [PubMed]
  294. Tsuchiya, T.; Nakayama, T.; Ariga, K. Nanoarchitectonics intelligence with atomic switch and neuromorphic network system. Appl. Phys. Express 2022, 15, 100101. [Google Scholar] [CrossRef]
  295. Baek, S.; Kim, S.; Han, S.A.; Kim, Y.H.; Kim, S.; Kim, J.H. Synthesis strategies and nanoarchitectonics for high-performance transition metal dichalcogenide thin film field-effect transistors. ChemNanoMat 2023, 9, e202300104. [Google Scholar] [CrossRef]
  296. Zhang, H.; Lin, D.-Q.; Wang, Y.-C.; Li, Z.-X.; Hu, S.; Huang, L.; Zhang, X.-W.; Jin, D.; Sheng, C.-X.; Xu, C.-X.; et al. Hierarchical nanoarchitectonics of ultrathin 2D organic nanosheets for aqueous processed electroluminescent devices. Small 2023, 19, 2208174. [Google Scholar] [CrossRef] [PubMed]
  297. Azzaroni, O.; Piccinini, E.; Fenoy, G.; Marmisollé, W.; Ariga, K. Field-effect transistors engineered via solution-based layer-by-layer nanoarchitectonics. Nanotechnology 2023, 34, 472001. [Google Scholar] [CrossRef]
  298. Halder, S.; Chakraborty, C. Fe(II)–Pt(II) based metallo-supramolecular macrocycle nanoarchitectonics for high-performance gel-state electrochromic device. Dye. Pigment. 2023, 212, 111131. [Google Scholar] [CrossRef]
  299. Zhou, F.; Zhao, Y.; Fu, F.; Liu, L.; Luo, Z. Thickness nanoarchitectonics with edge-enhanced Raman, polarization Raman, optoelectronic properties of GaS nanosheets devices. Crystals 2023, 13, 1506. [Google Scholar] [CrossRef]
  300. Chen, G.; Singh, S.K.; Takeyasu, K.; Hill, J.P.; Nakamura, J.; Ariga, K. Versatile nanoarchitectonics of Pt with morphology control of oxygen reduction reaction catalysts. Sci. Technol. Adv. Mater. 2022, 23, 413–423. [Google Scholar] [CrossRef]
  301. Tang, Y.; Yang, C.; Xu, X.; Kang, Y.; Henzie, J.; Que, W.; Yamauchi, Y. MXene nanoarchitectonics: Defect-engineered 2D MXenes towards enhanced electrochemical water splitting. Adv. Energy Mater. 2022, 12, 2103867. [Google Scholar] [CrossRef]
  302. Liu, X.; Chen, T.; Xue, Y.; Fan, J.; Shen, S.; Hossain, M.S.A.; Amin, M.A.; Pan, L.; Xu, X.; Yamauchi, Y. Nanoarchitectonics of MXene/semiconductor heterojunctions toward artificial photosynthesis via photocatalytic CO2 reduction. Coord. Chem. Rev. 2022, 459, 214440. [Google Scholar] [CrossRef]
  303. Feng, J.-C.; Xia, H. Application of nanoarchitectonics in moist-electric generation. Beilstein J. Nanotechnol. 2022, 13, 1185–1200. [Google Scholar] [CrossRef] [PubMed]
  304. Zhang, X.; Yang, P. g-C3N4 Nanosheet nanoarchitectonics: H2 Generation and CO2 reduction. ChemNanoMat 2023, 9, e202300041. [Google Scholar] [CrossRef]
  305. Ravipati, M.; Badhulika, S. Solvothermal synthesis of hybrid nanoarchitectonics nickel-metal organic framework modified nickel foam as a bifunctional electrocatalyst for direct urea and nitrate fuel cell. Adv. Powder Technol. 2023, 34, 104087. [Google Scholar] [CrossRef]
  306. Zhang, X.; Matras-Postolek, K.; Yang, P.; Jiang, S.P. Z-scheme WOx/Cu-g-C3N4 heterojunction nanoarchitectonics with promoted charge separation and transfer towards efficient full solar-spectrum photocatalysis. J. Colloid Interface Sci. 2023, 636, 646–656. [Google Scholar] [CrossRef]
  307. Kumar, A.V.N.; Yin, S.; Wang, Z.; Qian, X.; Yang, D.; Xu, Y.; Li, X.; Wang, H.; Wang, L. Direct fabrication of bimetallic AuPt nanobrick spherical nanoarchitectonics for the oxygen reduction reaction. New J. Chem. 2019, 43, 9628–9633. [Google Scholar] [CrossRef]
  308. Kim, J.; Kim, J.H.; Ariga, K. Redox-active polymers for energy storage nanoarchitectonics. Joule 2017, 1, 739–768. [Google Scholar] [CrossRef]
  309. Ezika, A.C.; Sadiku, E.R.; Ray, S.S.; Hamam, Y.; Folorunso, O.; Adekoya, C.J. Emerging advancements in polypyrrole MXene hybrid nanoarchitectonics for capacitive energy storage applications. J. Inorg. Organomet. Polym. 2022, 32, 1521–1540. [Google Scholar] [CrossRef]
  310. Yan, D.; Liu, L.; Wang, X.; Xu, K.; Zhong, J. Biomass-derived activated carbon nanoarchitectonics with hibiscus flowers for high-performance supercapacitor electrode applications. Chem. Eng. Technol. 2022, 45, 649–657. [Google Scholar] [CrossRef]
  311. Ramadass, K.; Sathish, C.; Singh, G.; Ruban, S.M.; Ruban, A.M.; Bahadur, R.; Kothandam, G.; Belperio, T.; Marsh, J.; Karakoti, A.; et al. Morphologically tunable nanoarchitectonics of mixed kaolin-halloysite derived nitrogen-doped activated nanoporous carbons for supercapacitor and CO2 capture applications. Carbon 2022, 192, 133–144. [Google Scholar] [CrossRef]
  312. Olivares, R.D.O.; Lobato-Peralta, D.R.; Arias, D.M.; Okolie, J.A.; Cuentas-Gallegos, A.K.; Sebastian, P.J.; Mayer, A.R.; Okoye, P.U. Production of nanoarchitectonics corncob activated carbon as electrode material for enhanced supercapacitor performance. J. Energy Storage 2022, 55, 105447. [Google Scholar] [CrossRef]
  313. Shinde, P.A.; Chodankar, N.R.; Kim, H.-J.; Abdelkareem, M.A.; Ghaferi, A.A.; Han, Y.K.; Olabi, A.G.; Ariga, K. Ultrastable 1T-2H WS2 heterostructures by nanoarchitectonics of phosphorus-triggered phase transition for hybrid supercapacitors. ACS Energy Lett. 2023, 8, 4474–4487. [Google Scholar] [CrossRef]
  314. Liu, S.; Wei, C.; Wang, H.; Yang, W.; Zhang, J.; Wang, Z.; Zhao, W.; Lee, P.S.; Ai, G. Processable nanoarchitectonics of two-dimensional metallo-supramolecular polymer for electrochromic energy storage devices with high coloration efficiency and stability. Nano Energy 2023, 110, 108337. [Google Scholar] [CrossRef]
  315. Li, T.; Dong, H.; Shi, Z.; Yue, H.; Yin, Y.; Li, X.; Zhang, H.; Wu, X.; Li, B.; Yang, S. Composite nanoarchitectonics with CoS2 nanoparticles embedded in graphene sheets for an anode for lithium-ion batteries. Nanomaterials 2022, 12, 724. [Google Scholar] [CrossRef] [PubMed]
  316. Kalidass, J.; Anandan, S.; Sivasankar, T. Sonoelectrochemical nanoarchitectonics of crystalline mesoporous magnetite @ manganese oxide nanocomposite as an alternate anode material for energy-storage applications. Crystals 2023, 13, 557. [Google Scholar] [CrossRef]
  317. Shin, M.; Awasthi, G.P.; Sharma, K.P.; Pandey, P.; Park, M.; Ojha, G.P.; Yu, C. Nanoarchitectonics of three-dimensional carbon nanofiber-supported hollow copper sulfide spheres for asymmetric supercapacitor applications. Int. J. Mol. Sci. 2023, 24, 9685. [Google Scholar] [CrossRef]
  318. Ali, N.; Funmilayo, O.R.; Khan, A.; Ali, F.; Bilal, M.; Yang, Y.; Akhter, M.S.; Zhou, C.; Wenjie, Y.; Iqbal, H.M.N. Nanoarchitectonics: Porous hydrogel as bio-sorbent for effective remediation of hazardous contaminants. J. Inorg. Organomet. Polym. 2022, 32, 3301–3320. [Google Scholar] [CrossRef]
  319. Ismail, K.S.I.K.; Tajudin, A.A.; Ikeno, S.; Hamzah, A.S.A. Heteroligand nanoarchitectonics of functionalized gold nanoparticle for Hg2+ detection. J. Nanopart. Res. 2022, 24, 253. [Google Scholar] [CrossRef]
  320. Salehipour, M.; Rezaei, S.; Asadi Khalili, H.F.; Motaharian, A.; Manzari, M.M. Nanoarchitectonics of enzyme/metal–organic framework composites for wastewater treatment. J. Inorg. Organomet. Polym. 2022, 32, 3321–3338. [Google Scholar] [CrossRef]
  321. Ren, Z.; Yang, X.; Ye, B.; Zhang, W.; Zhao, Z. Biomass-derived mesoporous nanoarchitectonics with magnetic MoS2 and activated carbon for enhanced adsorption of industrial cationic dye and tetracycline contaminants. Nano 2022, 17, 2250085. [Google Scholar] [CrossRef]
  322. Maimaitizi, H.; Abulizi, A.; Talifu, D.; Tursun, Y. Nanoarchitectonics of chlorophyll and Mg co-modified hierarchical BiOCl microsphere as an efficient photocatalyst for CO2 reduction and ciprofloxacin degradation. Adv. Powder Technol. 2022, 33, 103562. [Google Scholar] [CrossRef]
  323. Bhadra, B.N.; Shrestha, L.K.; Ariga, K. Porous carbon nanoarchitectonics for the environment: Detection and adsorption. CrystEngComm 2022, 24, 6804–6824. [Google Scholar] [CrossRef]
  324. Barreca, D.; Maccato, C. Nanoarchitectonics of metal oxide materials for sustainable technologies and environmental applications. CrystEngComm 2023, 25, 3968–3987. [Google Scholar] [CrossRef]
  325. Deng, G.; Xie, L.; Xu, S.; Kang, X.; Ma, J. Fiber nanoarchitectonics for pre-treatments in facile detection of short-chain fatty acids in waste water and faecal samples. Polymers 2021, 13, 3906. [Google Scholar] [CrossRef]
  326. Si, R.; Chen, Y.; Wang, D.; Yu, D.; Ding, Q.; Li, R.; Wu, C. Nanoarchitectonics for high adsorption capacity carboxymethyl cellulose nanofibrils-based adsorbents for efficient Cu2+ removal. Nanomaterials 2022, 12, 160. [Google Scholar] [CrossRef]
  327. Ashraf, I.; Li, R.; Chen, B.; Al-Ansari, N.; Aslam, M.R.; Altaf, A.R.; Elbeltagi, A. Nanoarchitectonics and kinetics insights into fluoride removal from drinking water using magnetic tea biochar. Int. J. Environ. Res. Public Health 2022, 19, 13092. [Google Scholar] [CrossRef]
  328. Shao, Y.; Xiang, L.; Zhang, W.; Chen, Y. Responsive shape-shifting nanoarchitectonics and its application in tumor diagnosis and therapy. J. Control. Release 2022, 352, 600–618. [Google Scholar] [CrossRef]
  329. Kumbhar, P.; Kolekar, K.; Khot, C.; Dabhole, S.; Salawi, A.; Sabei, F.A.; Mohite, A.; Kole, K.; Mhatre, S.; Jha, N.K.; et al. Co-crystal nanoarchitectonics as an emerging strategy in attenuating cancer: Fundamentals and applications. J. Control. Release 2023, 353, 1150–1170. [Google Scholar] [CrossRef]
  330. Tian, B.; Liu, J.; Guo, S.; Li, A.; Wan, J.-B. Macromolecule-based hydrogels nanoarchitectonics with mesenchymal stem cells for regenerative medicine: A review. Int. J. Biol. Macromol. 2023, 243, 125161. [Google Scholar] [CrossRef]
  331. Aziz, T.; Nadeem, A.A.; Sarwar, A.; Perveen, I.; Hussain, N.; Khan, A.A.; Daudzai, Z.; Cui, H.; Lin, L. Particle nanoarchitectonics for nanomedicine and nanotherapeutic drugs with special emphasis on nasal drugs and aging. Biomedicines 2023, 11, 354. [Google Scholar] [CrossRef] [PubMed]
  332. Sutrisno, L.; Ariga, K. Pore-engineered nanoarchitectonics for cancer therapy. NPG Asia Mater. 2023, 15, 21. [Google Scholar] [CrossRef]
  333. Li, B.; Huang, Y.; Zou, Q. Peptide-based nanoarchitectonics for the treatment of liver fibrosis. ChemBioChem 2023, 24, e202300002. [Google Scholar] [CrossRef] [PubMed]
  334. Kim, S.; Baek, S.; Sluyter, R.; Konstantinov, K.; Kim, J.H.; Kim, S.; Kim, Y.H. Wearable and implantable bioelectronics as eco-friendly and patient-friendly integrated nanoarchitectonics for next-generation smart healthcare technology. EcoMat 2023, 5, e12356. [Google Scholar] [CrossRef]
  335. Wang, Y.-M.; Xu, Y.; Zhang, X.; Cui, Y.; Liang, Q.; Liu, C.; Wang, X.; Wu, S.; Yang, R. Single Nano-sized metal–organic framework for bio-nanoarchitectonics with in vivo fluorescence imaging and chemo-photodynamic therapy. Nanomaterials 2022, 12, 287. [Google Scholar] [CrossRef] [PubMed]
  336. Yuan, Y.; Chen, L.; Shi, Z.; Chen, J. Micro/nanoarchitectonics of 3D printed scaffolds with excellent biocompatibility prepared using femtosecond laser two-photon polymerization for tissue engineering applications. Nanomaterials 2022, 12, 391. [Google Scholar] [CrossRef] [PubMed]
  337. Carrara, S.; Rouvier, F.; Auditto, S.; Brunel, F.; Jeanneau, C.; Camplo, M.; Sergent, M.; About, I.; Bolla, J.-M.; Raimundo, J.-M. Nanoarchitectonics of electrically activable phosphonium self-assembled monolayers to efficiently kill and tackle bacterial infections on demand. Int. J. Mol. Sci. 2022, 23, 2183. [Google Scholar] [CrossRef]
  338. Kulikov, O.A.; Zharkov, M.N.; Ageev, V.P.; Yakobson, D.E.; Shlyapkina, V.I.; Zaborovskiy, A.V.; Inchina, V.I.; Balykova, L.A.; Tishin, A.M.; Sukhorukov, G.B.; et al. Magnetic Hyperthermia Nanoarchitectonics via iron oxide nanoparticles stabilised by oleic acid: Anti-tumour efficiency and safety evaluation in animals with transplanted carcinoma. Int. J. Mol. Sci. 2022, 23, 4234. [Google Scholar] [CrossRef]
  339. Do, T.T.A.; Wicaksono, K.; Soendoro, A.; Imae, T.; Garcia-Celma, M.J.; Grijalvo, S. Complexation nanoarchitectonics of carbon dots with doxorubicin toward photodynamic anti-cancer therapy. J. Funct. Biomater. 2022, 13, 219. [Google Scholar] [CrossRef]
  340. Jayachandran, P.; Ilango, S.; Suseela, V.; Nirmaladevi, R.; Shaik, M.R.; Khan, M.; Khan, M.; Shaik, B. Green synthesized silver nanoparticle-loaded liposome-based nanoarchitectonics for cancer management: In Vitro drug release analysis. Biomedicines 2023, 11, 217. [Google Scholar] [CrossRef]
  341. Papadopoulou-Fermeli, N.; Lagopati, N.; Pippa, N.; Sakellis, E.; Boukos, N.; Gorgoulis, V.G.; Gazouli, M.; Pavlatou, E.A. Composite nanoarchitectonics of photoactivated titania-based materials with anticancer properties. Pharmaceutics 2023, 15, 135. [Google Scholar] [CrossRef]
  342. Osada, M.; Sasaki, T. Nanoarchitectonics in dielectric/ferroelectric layered perovskites: From bulk 3D systems to 2D nanosheets. Dalton Trans. 2018, 47, 2841–2851. [Google Scholar] [CrossRef] [PubMed]
  343. Vranckx, C.; Lambricht, L.; Préat, V.; Cornu, O.; Dupont-Gillain, C.; van der Straeten, A. Layer-by-layer nanoarchitectonics using protein–polyelectrolyte complexes toward a generalizable tool for protein surface immobilization. Langmuir 2022, 38, 5579–5589. [Google Scholar] [CrossRef] [PubMed]
  344. Ariga, K. Chemistry of materials nanoarchitectonics for two-dimensional films: Langmuir–Blodgett, layer-by-layer assembly, and newcomers. Chem. Mater. 2023, 35, 5233–5254. [Google Scholar] [CrossRef]
  345. Banszerus, L.; Schmitz, M.; Engels, S.; Dauber, J.; Oellers, M.; Haupt, F.; Watanabe, K.; Taniguchi, T.; Beschoten, B.; Stampfer, C. Ultrahigh-mobility graphene devices from chemical vapor deposition on reusable copper. Sci. Adv. 2015, 1, e1500222. [Google Scholar] [CrossRef]
  346. Zhou, X.; Gan, L.; Tian, W.; Zhang, Q.; Jin, S.; Li, H.; Bando, Y.; Golberg, D.; Zhai, T. Ultrathin SnSe2 flakes grown by chemical vapor deposition for high-performance photodetectors. Adv. Mater. 2015, 27, 8035–8041. [Google Scholar] [CrossRef] [PubMed]
  347. Kaneko, K.; Uno, K.; Jinno, R.; Fujita, S. Prospects for phase engineering of semi-stable Ga2O3 semiconductor thin films using mist chemical vapor deposition. J. Appl. Phys. 2022, 131, 090902. [Google Scholar] [CrossRef]
  348. Murata, T.; Minami, K.; Yamazaki, T.; Sato, T.; Koinuma, H.; Ariga, K.; Matsuki, N. Nanometer-flat DNA-featured thin films prepared via laser molecular beam deposition under high-vacuum for selective methanol sensing. Bull. Chem. Soc. Jpn. 2023, 96, 29–34. [Google Scholar] [CrossRef]
  349. Decher, G. Fuzzy nanoassemblies: Toward layered polymeric multicomposites. Science 1997, 277, 1232–1237. [Google Scholar] [CrossRef]
  350. Rydzek, G.; Ji, Q.; Li, M.; Schaaf, P.; Hill, J.P.; Boulmedais, F.; Ariga, K. Electrochemical nanoarchitectonics and layer-by-layer assembly: From basics to future. Nano Today 2015, 10, 138–167. [Google Scholar] [CrossRef]
  351. Richardson, J.J.; Björnmalm, M.; Caruso, F. Technology-driven layer-by-layer assembly of nanofilms. Science 2015, 348, aaa2491. [Google Scholar] [CrossRef] [PubMed]
  352. Wang, C.; Park, M.J.; Yu, H.; Matsuyama, H.; Drioli, E.; Shon, H.K. Recent advances of nanocomposite membranes using layer-by-layer assembly. J. Membr. Sci. 2022, 661, 120926. [Google Scholar] [CrossRef]
  353. Ariga, K.; Lvov, Y.; Decher, G. There is still plenty of room for layer-by-layer assembly for constructing nanoarchitectonics-based materials and devices. Phys. Chem. Chem. Phys. 2022, 24, 4097–4115. [Google Scholar] [CrossRef] [PubMed]
  354. Yoon, J.-W.; Kim, J.-S.; Kim, T.-H.; Hong, Y.J.; Kang, Y.C.; Lee, J.-H. A new strategy for humidity independent oxide chemiresistors: Dynamic self-refreshing of In2O3 sensing surface assisted by layer-by-layer coated CeO2 nanoclusters. Small 2016, 12, 4229–4240. [Google Scholar] [CrossRef] [PubMed]
  355. Yang, L.; Li, L.; Li, H.; Wang, T.; Ren, X.; Cheng, Y.; Li, Y.; Huang, Q. Layer-by-layer assembled smart antibacterial coatings via mussel-inspired polymerization and dynamic covalent chemistry. Adv. Healthc. Mater. 2022, 11, 2200112. [Google Scholar] [CrossRef] [PubMed]
  356. Lee, T.; Ohshiro, K.; Watanabe, T.; Hyeon-Deuk, K.; Kim, D. Temperature-dependent exciton dynamics in CdTe quantum dot superlattices fabricated via layer-by-layer assembly. Adv. Opt. Mater. 2022, 10, 2102781. [Google Scholar] [CrossRef]
  357. Kaganer, V.M.; Möhwald, H.; Dutta, P. Structure and phase transitions in Langmuir monolayers. Rev. Mod. Phys. 1999, 71, 779–819. [Google Scholar] [CrossRef]
  358. Ariga, K.; Yamauchi, Y.; Mori, T.; Hill, J.P. 25th Anniversary article: What can be done with the Langmuir-Blodgett method? Recent developments and its critical role in materials science. Adv. Mater. 2013, 25, 6477–6512. [Google Scholar] [CrossRef]
  359. Ariga, K. Don’t forget Langmuir–Blodgett films 2020: Interfacial nanoarchitectonics with molecules, materials, and living objects. Langmuir 2020, 36, 7158–7180. [Google Scholar] [CrossRef]
  360. Li, X.; Zhang, G.; Bai, X.; Sun, X.; Wang, X.; Wang, E.; Dai, H. Highly conducting graphene sheets and Langmuir–Blodgett films. Nat. Nanotechnol. 2008, 3, 538–542. [Google Scholar] [CrossRef]
  361. Fang, C.; Yoon, I.; Hubble, D.; Tran, T.-N.; Kostecki, R.; Liu, G. Recent applications of Langmuir–Blodgett technique in battery research. ACS Appl. Mater. Interfaces 2022, 14, 2431–2439. [Google Scholar] [CrossRef] [PubMed]
  362. Oliveira, O.N., Jr.; Caseli, L.; Ariga, K. The Past and the future of Langmuir and Langmuir–Blodgett films. Chem. Rev. 2022, 122, 6459–6513. [Google Scholar] [CrossRef] [PubMed]
  363. Krishnan, V.; Kasuya, Y.; Ji, Q.; Sathish, M.; Shrestha, L.K.; Ishihara, S.; Minami, K.; Morita, H.; Yamazaki, T.; Hanagata, N.; et al. Vortex-aligned fullerene nanowhiskers as a scaffold for orienting cell growth. ACS Appl. Mater. Interfaces 2015, 7, 15667–15673. [Google Scholar] [CrossRef] [PubMed]
  364. Ito, M.; Yamashita, Y.; Tsuneda, Y.; Mori, T.; Takeya, J.; Watanabe, S.; Ariga, K. 100 °C-Langmuir–Blodgett method for fabricating highly oriented, ultrathin films of polymeric semiconductors. ACS Appl. Mater. Interfaces 2020, 12, 56522–56529. [Google Scholar] [CrossRef] [PubMed]
  365. Ito, M.; Yamashita, Y.; Mori, T.; Chiba, M.; Futae, T.; Takeya, J.; Watanabe, S.; Ariga, K. Hyper 100 °C Langmuir–Blodgett (Langmuir–Schaefer) technique for organized ultrathin film of polymeric semiconductors. Langmuir 2022, 38, 5237–5247. [Google Scholar] [CrossRef]
  366. Ariga, K. Materials nanoarchitectonics in a two-dimensional world within a nanoscale distance from the liquid phase. Nanoscale 2022, 14, 10610–10629. [Google Scholar] [CrossRef]
  367. Ariga, K. Liquid–liquid interfacial nanoarchitectonics. Small 2023, 2305636. [Google Scholar] [CrossRef]
  368. Negi, S.; Hamori, M.; Kitagishi, H.; Kano, K. Highly ordered monolayers of an optically active amphiphilic pyrene derivative at the air–water interface. Bull. Chem. Soc. Jpn. 2022, 95, 1537–1545. [Google Scholar] [CrossRef]
  369. Negi, S.; Hamori, M.; Kubo, Y.; Kitagishi, H.; Kano, K. Monolayer formation and chiral recognition of binaphthyl amphiphiles at the air–water interface. Bull. Chem. Soc. Jpn. 2023, 96, 48–56. [Google Scholar] [CrossRef]
  370. Takase, S.; Aritsu, T.; Sakamoto, Y.; Sakuno, Y.; Shimizu, Y. Preparation of highly conductive phthalocyaninato-cobalt iodide at the interface between aqueous KI solution and organic solvent and catalytic properties for electrochemical reduction of CO2. Bull. Chem. Soc. Jpn. 2023, 96, 649–653. [Google Scholar] [CrossRef]
  371. Ariga, K.; Mori, T.; Ishihara, S.; Kawakami, K.; Hill, J.P. Bridging the difference to the billionth-of-a-meter length scale: How to operate nanoscopic machines and nanomaterials by using macroscopic actions. Chem. Mater. 2014, 26, 519–532. [Google Scholar] [CrossRef]
  372. Adachi, J.; Naito, M.; Sugiura, S.; Le, N.H.-T.; Nishimura, S.; Huang, S.; Suzuki, S.; Kawamorita, S.; Komiya, N.; Hill, J.P.; et al. Coordination amphiphile: Design of planar-coordinated platinum complexes for monolayer formation at an air-water interface based on ligand characteristics and molecular topology. Bull. Chem. Soc. Jpn. 2022, 95, 889–897. [Google Scholar] [CrossRef]
  373. Ariga, K.; Mori, T.; Hill, J.P. Mechanical control of nanomaterials and nanosystems. Adv. Mater. 2012, 24, 158–176. [Google Scholar] [CrossRef] [PubMed]
  374. Ariga, K. Molecular machines and microrobots: Nanoarchitectonics developments and on-water performances. Micromachines 2023, 14, 25. [Google Scholar] [CrossRef] [PubMed]
  375. Adachi, J.; Mori, T.; Inoue, R.; Naito, M.; Ngoc Le, N.H.-T.; Kawamorita, S.; Hill, J.P.; Naota, T.; Ariga, K. Emission control by molecular manipulation of double-paddled binuclear PtII complexes at the air-water interface. Chem. Asian J. 2020, 15, 406–414. [Google Scholar] [CrossRef] [PubMed]
  376. Maeda, T.; Mori, T.; Ikeshita, M.; Ma, S.C.; Muller, G.; Ariga, K.; Naota, T. Vortex flow-controlled circularly polarized luminescence of achiral Pt(II) complex aggregates assembled at the air-water interface. Small Methods 2022, 6, 2200936. [Google Scholar] [CrossRef] [PubMed]
  377. Seong, H.-G.; Fink, Z.; Chen, Z.; Emrick, T.; Russell, T.P. Bottlebrush polymers at liquid interfaces: Assembly dynamics, mechanical properties, and all-liquid printed constructs. ACS Nano 2023, 17, 14731–14741. [Google Scholar] [CrossRef] [PubMed]
  378. Flanagan, J.C.; Valentine, M.J.; Baiz, C.R. Ultrafast dynamics at lipid–water interfaces. Acc. Chem. Res. 2020, 53, 1860–1868. [Google Scholar] [CrossRef]
  379. Hammel, M.; Tainer, J.A. X-ray scattering reveals disordered linkers and dynamic interfaces in complexes and mechanisms for DNA double-strand break repair impacting cell and cancer biology. Protein Sci. 2021, 30, 1735–1756. [Google Scholar] [CrossRef]
  380. Hosseinpour, S.; Roeters, S.J.; Bonn, M.; Peukert, W.; Woutersen, S.; Weidner, T. Structure and dynamics of interfacial peptides and proteins from vibrational sum-frequency generation spectroscopy. Chem. Rev. 2020, 120, 3420–3465. [Google Scholar] [CrossRef]
  381. Agashe, C.; Varshney, R.; Sangwan, R.; Gill, A.K.; Alam, M.; Patra, D. Anisotropic compartmentalization of the liquid–liquid interface using dynamic imine chemistry. Langmuir 2022, 38, 8296–8303. [Google Scholar] [CrossRef] [PubMed]
  382. Dai, W.; Shao, F.; Szczerbiński, J.; McCaffrey, R.; Zenobi, R.; Jin, Y.; Schlüter, A.D.; Zhang, W. Synthesis of a two-dimensional covalent organic monolayer through dynamic imine chemistry at the air/water interface. Angew. Chem. Int. Ed. 2016, 55, 213–217. [Google Scholar] [CrossRef] [PubMed]
  383. Makiura, R.; Motoyama, S.; Umemura, Y.; Yamanaka, H.; Sakata, O.; Kitagawa, H. Surface nano-architecture of a metal–organic framework. Nat. Mater. 2010, 9, 565–571. [Google Scholar] [CrossRef] [PubMed]
  384. Kambe, T.; Sakamoto, R.; Kusamoto, T.; Pal, T.; Fukui, N.; Hoshiko, K.; Shimojima, T.; Wang, Z.; Hirahara, T.; Ishizaka, K.; et al. Redox control and high conductivity of nickel bis(dithiolene) complex p-nanosheet: A potential organic two-dimensional topological insulator. J. Am. Chem. Soc. 2014, 136, 14357–14360. [Google Scholar] [CrossRef] [PubMed]
  385. Dey, K.; Pal, M.; Rou, K.C.; Kunjattu, H.S.; Das, A.; Mukherjee, R.; Kharul, U.K.; Banerjee, R. Selective molecular separation by interfacially crystallized covalent organic framework thin films. J. Am. Chem. Soc. 2017, 139, 13083–13091. [Google Scholar] [CrossRef] [PubMed]
  386. Sun, K.; Wang, C.; Dong, Y.; Guo, P.; Cheng, P.; Fu, Y.; Liu, D.; He, D.; Das, S.; Negishi, Y. Ion-selective covalent organic framework membranes as a catalytic polysulfide trap to arrest the redox shuttle effect in lithium–sulfur batteries. ACS Appl. Mater. Interfaces 2022, 14, 4079–4090. [Google Scholar] [CrossRef] [PubMed]
  387. Du, J.; Sun, Q.; He, W.; Liu, L.; Song, Z.; Yao, A.; Ma, J.; Cao, D.; Hassan, S.U.; Guan, J.; et al. A 2D soft covalent organic framework membrane prepared via a molecular bridge. Adv. Mater. 2023, 35, 2300975. [Google Scholar] [CrossRef]
  388. Ohata, T.; Tachimoto, K.; Takeno, K.J.; Nomoto, A.; Watanabe, T.; Hirosawa, I.; Makiura, R. Influence of the solvent on the assembly of Ni3(hexaiminotriphenylene)2 metal–organic framework nanosheets at the air/liquid interface. Bull. Chem. Soc. Jpn. 2023, 96, 274–282. [Google Scholar] [CrossRef]
  389. Bae, S.; Kim, H.; Lee, Y.; Xu, X.; Park, J.-S.; Zheng, Y.; Balakrishnan, J.; Lei, T.; Kim, H.R.; Song, Y.I.; et al. Roll-to-roll production of 30-inch graphene films for transparent electrodes. Nat. Nanotecnol. 2010, 5, 574–578. [Google Scholar] [CrossRef]
  390. Song, J.; Murata, T.; Tsai, K.-C.; Jia, X.; Sciortino, F.; Ma, R.; Yamauchi, Y.; Hill, J.P.; Shrestha, L.K.; Ariga, K. Fullerphene nanosheets: A bottom-Up 2D material for single-carbon-atom-level molecular discrimination. Adv. Mater. Interfaces 2022, 9, 2102241. [Google Scholar] [CrossRef]
  391. Fu, M.; Chen, W.; Lei, Y.; Yu, H.; Lin, Y.; Terrones, M. Biomimetic construction of ferrite quantum dot/graphene heterostructure for enhancing ion/charge transfer in supercapacitors. Adv. Mater. 2023, 35, 2300940. [Google Scholar] [CrossRef] [PubMed]
  392. Mori, T.; Tanaka, H.; Dalui, A.; Mitoma, N.; Suzuki, K.; Matsumoto, M.; Aggarwal, N.; Patnaik, A.; Acharya, S.; Shrestha, L.K.; et al. Carbon nanosheets by morphology-retained carbonization of two-dimensional assembled anisotropic carbon nanorings. Angew. Chem. Int. Ed. 2018, 57, 9679–9683. [Google Scholar] [CrossRef] [PubMed]
  393. Du, X.; Zhang, L.; Chen, R.; You, J.; Ma, Y.; Wang, J.; Wu, Y.; Liu, B.; Zhao, K.; Chen, J.; et al. Spontaneous interface healing by a dynamic liquid-crystal transition for high-performance perovskite solar cells. Adv. Mater. 2022, 34, 2207362. [Google Scholar] [CrossRef] [PubMed]
  394. Yoo, E.; Kim, J.; Hosono†, E.; Zhou, H.-S.; Kudo, T.; Honma, I. Large reversible Li storage of graphene nanosheet families for use in rechargeable lithium ion batteries. Nano Lett. 2008, 8, 2277–2282. [Google Scholar] [CrossRef] [PubMed]
  395. Yoshino, A. The lithium-ion battery: Two breakthroughs in development and two reasons for the Nobel Prize. Bull. Chem. Soc. Jpn. 2022, 95, 195–197. [Google Scholar] [CrossRef]
  396. Kondo, Y.; Abe, T.; Yamada, Y. Kinetics of interfacial ion transfer in lithium-ion batteries: Mechanism understanding and improvement strategies. ACS Appl. Mater. Interfaces 2022, 14, 22706–22718. [Google Scholar] [CrossRef]
  397. Qiu, Y.-F.; Murayama, H.; Fujitomo, C.; Kawai, S.; Haruta, A.; Hiasa, T.; Mita, H.; Motohashi, K.; Yamamoto, E.; Tokunaga, M. Oxidative decomposition mechanism of ethylene carbonate on positive electrodes in lithium-ion batteries. Bull. Chem. Soc. Jpn. 2023, 96, 444–451. [Google Scholar] [CrossRef]
  398. Matsumoto, F.; Yamada, M.; Tsuta, M.; Nakamura, S.; Ando, N.; Soma, N. Review of the structure and performance of through-holed anodes and cathodes prepared with a picosecond pulsed laser for lithium-ion batteries. Int. J. Extrem. Manuf. 2023, 5, 012001. [Google Scholar] [CrossRef]
  399. Komaba, S.; Hasegawa, T.; Dahbi, M.; Kubota, K. Potassium intercalation into graphite to realize high-voltage/high-power potassium-ion batteries and potassium-ion capacitors. Electrochem. Commun. 2015, 60, 172–175. [Google Scholar] [CrossRef]
  400. Hosaka, T.; Komaba, S. Development of nonaqueous electrolytes for high-voltage K-ion batteries. Bull. Chem. Soc. Jpn. 2022, 95, 569–581. [Google Scholar] [CrossRef]
  401. Kim, E.J.; Kumar, P.R.; Gossage, Z.T.; Kubota, K.; Hosaka, T.; Tatara, R.; Komaba, S. Active material and interphase structures governing performance in sodium and potassium ion batteries. Chem. Sci. 2022, 13, 6121–6158. [Google Scholar] [CrossRef] [PubMed]
  402. Nguyen, M.T.; Muramatsu, T.; Kheawhom, S.; Wattanakit, C.; Yonezawa, T. Impact of morphology and transition metal doping of vanadate nanowires without surface modification on the performance of aqueous zinc-ion batteries. Bull. Chem. Soc. Jpn. 2022, 95, 728–734. [Google Scholar] [CrossRef]
  403. Yang, W.; Yang, Y.; Yang, H.; Zhou, H. Regulating water activity for rechargeable zinc-ion batteries: Progress and perspective. ACS Energy Lett. 2022, 7, 2515–2530. [Google Scholar] [CrossRef]
  404. Ganesan, P.; Ishihara, A.; Staykov, A.; Nakashima, N. Recent advances in nanocarbon-based nonprecious metal catalysts for oxygen/hydrogen reduction/evolution reactions and Zn-air battery. Bull. Chem. Soc. Jpn. 2023, 96, 429–443. [Google Scholar] [CrossRef]
  405. Gopalakrishnan, M.; Ganesan, S.; Nguyen, M.T.; Yonezawa, T.; Praserthdam, S.; Pornprasertsuk, R.; Kheawhom, S. Critical roles of metal–organic frameworks in improving the Zn anode in aqueous zinc-ion batteries. Chem. Eng. J. 2023, 457, 141334. [Google Scholar] [CrossRef]
  406. Guo, Z.; Fan, L.; Zhao, C.; Chen, A.; Liu, N.; Zhang, Y.; Zhang, N. A dynamic and self-adapting interface coating for stable Zn-metal anodes. Adv. Mater. 2022, 34, 2105133. [Google Scholar] [CrossRef]
  407. Miyazawa, K. Synthesis of fullerene nanowhiskers using the liquid–liquid interfacial precipitation method and their mechanical, electrical and superconducting properties. Sci. Technol. Adv. Mater. 2015, 16, 013502. [Google Scholar] [CrossRef]
  408. Miyazawa, K.; Kuwasaki, Y.; Obayashi, A.; Kuwabara, M. C60 nanowhiskers formed by the liquid–liquid interfacial precipitation method. J. Mater. Res. 2002, 17, 83–88. [Google Scholar] [CrossRef]
  409. Koya, I.; Yokoyama, Y.; Sakka, T.; Nishi, N. Formation of Au nanofiber/fullerene nanowhisker 1D/1D composites via reductive deposition at the interface between an ionic liquid and water. Chem. Lett. 2022, 51, 643–645. [Google Scholar] [CrossRef]
  410. Sathish, M.; Miyazawa, K. Size-tunable hexagonal fullerene (C60) nanosheets at the liquid−liquid interface. J. Am. Chem. Soc. 2007, 129, 13816–13817. [Google Scholar] [CrossRef]
  411. Chen, G.; Shrestha, L.K.; Ariga, K. Zero-to-two nanoarchitectonics: Fabrication of two-dimensional materials from zero-dimensional fullerene. Molecules 2021, 26, 4636. [Google Scholar] [CrossRef] [PubMed]
  412. Park, C.; Yoon, E.; Kawano, M.; Joo, T.; Choi, H.C. Self-crystallization of C70 cubes and remarkable enhancement of photoluminescence. Angew. Chem. Int. Ed. 2010, 49, 9670–9675. [Google Scholar] [CrossRef] [PubMed]
  413. Kim, J.; Park, C.; Choi, H.C. Selective growth of a C70 crystal in a mixed solvent system: From cube to tube. Chem. Mater. 2015, 27, 2408–2413. [Google Scholar] [CrossRef]
  414. Maji, S.; Shrestha, L.K.; Ariga, K. Nanoarchitectonics for hierarchical fullerene nanomaterials. Nanomaterials 2021, 11, 2146. [Google Scholar] [CrossRef] [PubMed]
  415. Zhang, W.-M.; Hu, J.-S.; Guo, Y.-G.; Zheng, S.-F.; Zhong, L.-S.; Song, W.-G.; Wan, L.-J. Tin-nanoparticles encapsulated in elastic hollow carbon spheres for high-performance anode material in lithium-ion batteries. Adv. Mater. 2008, 20, 1160–1165. [Google Scholar] [CrossRef]
  416. Tan, H.; Li, Y.; Kim, J.; Takei, T.; Wang, Z.; Xu, X.; Wang, J.; Bando, Y.; Kang, Y.-M.; Tang, J.; et al. Sub-50 nm iron–nitrogen-doped hollow carbon sphere-encapsulated iron carbide nanoparticles as efficient oxygen reduction catalysts. Adv. Sci. 2018, 5, 1800120. [Google Scholar] [CrossRef] [PubMed]
  417. Yang, G.; Kuwahara, Y.; Mori, K.; Louis, C.; Yamashita, H. PdAg Alloy nanoparticles encapsulated in N-doped microporous hollow carbon spheres for hydrogenation of CO2 to formate. Appl. Catal. B Environ. 2021, 283, 119628. [Google Scholar] [CrossRef]
  418. Chen, G.; Sciortino, F.; Takeyasu, K.; Nakamura, J.; Hill, J.P.; Shrestha, L.K.; Ariga, K. Hollow spherical fullerene obtained by kinetically controlled liquid-liquid interfacial precipitation. Chem. Asian J. 2022, 17, e20220075. [Google Scholar] [CrossRef]
  419. Hsieh, C.-T.; Hsu, S.-h.; Maji, S.; Chahal, M.K.; Song, J.; Hill, J.P.; Ariga, K.; Shrestha, L.K. Post-assembly dimension-dependent face-selective etching of fullerene crystals. Mater. Horiz. 2020, 7, 787–795. [Google Scholar] [CrossRef]
  420. Chen, G.; Bhadra, B.N.; Sutrisno, L.; Shrestha, L.K.; Ariga, K. Fullerene rosette: Two-dimensional interactive nanoarchitectonics and selective vapor sensing. Int. J. Mol. Sci. 2022, 23, 5454. [Google Scholar] [CrossRef]
  421. Wei, Z.; Song, J.; Ma, R.; Ariga, K.; Shrestha, L.K. Self-assembled corn-husk-shaped fullerene crystals as excellent acid vapor sensors. Chemosensors 2022, 10, 16. [Google Scholar] [CrossRef]
  422. Tang, Q.; Bairi, P.; Shrestha, R.G.; Hill, J.P.; Ariga, K.; Zeng, H.; Ji, Q.; Shrestha, L.K. Quasi 2D mesoporous carbon microbelts derived from fullerene crystals as an electrode material for electrochemical supercapacitors. ACS Appl. Mater. Interfaces 2017, 9, 44458–44465. [Google Scholar] [CrossRef] [PubMed]
  423. Samukawa, S. Ultimate top-down etching processes for future nanoscale devices: Advanced neutral-beam etching. Jpn. J. Appl. Phys. 2006, 45, 2395. [Google Scholar] [CrossRef]
  424. Pennelli, G. Top down fabrication of long silicon nanowire devices by means of lateral oxidation. Microelectron. Eng. 2009, 86, 2139–2143. [Google Scholar] [CrossRef]
  425. Urs, K.M.B.; Sahoo, K.; Bhat, N.; Kamble, V. Complementary metal oxide semiconductor-compatible top-down fabrication of a Ni/NiO nanobeam room temperature hydrogen sensor device. ACS Appl. Electron. Mater. 2022, 4, 87–91. [Google Scholar] [CrossRef]
  426. Balzani, V.; Credi, A.; Venturi, M. The bottom-up approach to molecular-level devices and machines. Chem. Eur. J. 2022, 8, 5524–5532. [Google Scholar] [CrossRef]
  427. Li, C.; Huixia Feng, H.; Xu, H.; Chen, B.; Yang, T. An intelligent superhydrophilic/underwater superoleophobic temperature sensitive switch device with excellent targeted oil-water separation performance. Bull. Chem. Soc. Jpn. 2022, 95, 532–537. [Google Scholar] [CrossRef]
  428. Arjmand, T.; Legallais, M.; Nguyen, T.T.T.; Serre, P.; Vallejo-Perez, M.; Morisot, F.; Salem, B.; Ternon, C. Functional devices from bottom-up silicon nanowires: A Review. Nanomaterials 2022, 12, 1043. [Google Scholar] [CrossRef]
  429. Saito, Y.; Sasabe, H.; Tsuneyama, H.; Abe, S.; Matsuya, M.; Kawano, T.; Kori, Y.; Hanayama, T.; Kido, J. Quinoline-modified phenanthroline electron-transporters as n-type exciplex partners for highly efficient and stable deep-red OLEDs. Bull. Chem. Soc. Jpn. 2023, 96, 24–28. [Google Scholar] [CrossRef]
  430. Matsuya, M.; Sasabe, H.; Sumikoshi, S.; Hoshi, K.; Nakao, K.; Kumada, K.; Sugiyama, R.; Sato, R.; Kido, J. Highly luminescent aluminum complex with β-diketone ligands exhibiting near-unity photoluminescence quantum yield, thermally activated delayed fluorescence, and rapid radiative decay rate properties in solution-processed organic light-emitting devices. Bull. Chem. Soc. Jpn. 2023, 96, 183–189. [Google Scholar] [CrossRef]
  431. Ishii, M.; Yamashita, Y.; Watanabe, S.; Ariga, K.; Takeya, J. Doping of molecular semiconductors through proton-coupled electron transfer. Nature 2023, 622, 285–291. [Google Scholar] [CrossRef] [PubMed]
  432. Yamamoto, Y.; Nakano, S.; Shigeta, Y. Dynamical interaction analysis of proteins by a random forest-fragment molecular orbital (RF-FMO) method and application to Src tyrosine kinase. Bull. Chem. Soc. Jpn. 2023, 96, 42–47. [Google Scholar] [CrossRef]
  433. Mahmood, A.; Sandali, Y.; Wang, J.-L. Easy and fast prediction of green solvents for small molecule donor-based organic solar cells through machine learning. Phys. Chem. Chem. Phys. 2023, 25, 10417–10426. [Google Scholar] [CrossRef] [PubMed]
  434. Liang, Y.; Jiao, C.; Zhou, P.; Li, W.; Zang, Y.; Liu, Y.; Yang, G.; Liu, L.; Cheng, J.; Liang, G.; et al. Highly efficient perovskite solar cells with light management of surface antireflection. Bull. Chem. Soc. Jpn. 2023, 96, 148–155. [Google Scholar] [CrossRef]
  435. Yang, C.; Zhang, D.; Wang, D.; Luan, H.; Chen, X.; Yan, W. In Situ polymerized MXene/polypyrrole/hydroxyethyl cellulose-based flexible strain sensor enabled by machine learning for handwriting recognition. ACS Appl. Mater. Interfaces 2023, 15, 5811–5821. [Google Scholar] [CrossRef] [PubMed]
  436. Saito, N.; Nawachi, A.; Kondo, Y.; Choi, J.; Morimoto, H.; Ohshima, T. Functional group evaluation kit for digitalization of information on the functional group compatibility and chemoselectivity of organic reactions. Bull. Chem. Soc. Jpn. 2023, 96, 465–474. [Google Scholar] [CrossRef]
  437. Ramprasad, R.; Batra, R.; Pilania, G.; Mannodi-Kanakkithodi, A.; Kim, C. Machine learning in materials informatics: Recent applications and prospects. npj Comput. Mater. 2017, 3, 54. [Google Scholar] [CrossRef]
  438. Agrawala, A.; Choudhary, A. Perspective: Materials informatics and big data: Realization of the “fourth paradigm” of science in materials science. APL Mater. 2016, 4, 053208. [Google Scholar] [CrossRef]
  439. Hatakeyama-Sato, K.; Umeki, M.; Adachi, H.; Kuwata, N.; Hasegawa, G.; Oyaizu, K. Exploration of organic superionic glassy conductors by process and materials informatics with lossless graph database. npj Comput. Mater. 2022, 8, 170. [Google Scholar] [CrossRef]
  440. Chaikittisilp, W.; Yamauchi, Y.; Ariga, K. Material evolution with nanotechnology, nanoarchitectonics, and materials informatics: What will be the next paradigm shift in nanoporous materials? Adv. Mater. 2022, 34, 2107212. [Google Scholar] [CrossRef]
  441. Oviedo, L.R.; Oviedo, V.R.; Martins, M.O.; Fagan, S.B.; da Silva, W.L. Nanoarchitectonics: The role of artificial intelligence in the design and application of nanoarchitectures. J. Nanopart. Res. 2022, 24, 157. [Google Scholar] [CrossRef]
Figure 1. Outline of nanoarchitectonics, which combines nanotechnology with organic chemistry, inorganic chemistry, polymer chemistry, coordination chemistry, supramolecular chemistry, material chemistry, biochemistry, and microfabrication techniques.
Figure 1. Outline of nanoarchitectonics, which combines nanotechnology with organic chemistry, inorganic chemistry, polymer chemistry, coordination chemistry, supramolecular chemistry, material chemistry, biochemistry, and microfabrication techniques.
Materials 17 00271 g001
Figure 2. Two-dimensional nanoarchitectonics of binaphthyl molecules with an axial chirality in its racemic form (left) and its optically active S-isomer form (right). Reprinted with permission from [369]. Copyright 2023, Chemical Society of Japan.
Figure 2. Two-dimensional nanoarchitectonics of binaphthyl molecules with an axial chirality in its racemic form (left) and its optically active S-isomer form (right). Reprinted with permission from [369]. Copyright 2023, Chemical Society of Japan.
Materials 17 00271 g002
Figure 3. The nanoarchitectonics of the highly conductive charge transfer complex of (phthalocyaninato)cobalt iodide through a cobalt phthalocyanine crystal phase transition method by simply mixing a KI solution containing CF3COOH and cobalt phthalocyanine with a CH2Cl2 solution at the interface. Reprinted with permission from [370]. Copyright 2023, Chemical Society of Japan.
Figure 3. The nanoarchitectonics of the highly conductive charge transfer complex of (phthalocyaninato)cobalt iodide through a cobalt phthalocyanine crystal phase transition method by simply mixing a KI solution containing CF3COOH and cobalt phthalocyanine with a CH2Cl2 solution at the interface. Reprinted with permission from [370]. Copyright 2023, Chemical Society of Japan.
Materials 17 00271 g003
Figure 4. A new principle, submarine luminescence, as a luminescence control, involving molecular manipulation of double-paddle platinum complexes through compression at the air–water interface. Reprinted with permission from [375]. Copyright 2020, Wiley-VCH.
Figure 4. A new principle, submarine luminescence, as a luminescence control, involving molecular manipulation of double-paddle platinum complexes through compression at the air–water interface. Reprinted with permission from [375]. Copyright 2020, Wiley-VCH.
Materials 17 00271 g004
Figure 5. A flexible control of the intensity and signature of circularly polarized emissions from molecular aggregates of an achiral trans-bis(salicylaldiminato)platinum(II) complex upon precisely controlled clockwise and counterclockwise vortex motions at the air–water interface. Reprinted with permission from [376]. Copyright 2022, Wiley-VCH.
Figure 5. A flexible control of the intensity and signature of circularly polarized emissions from molecular aggregates of an achiral trans-bis(salicylaldiminato)platinum(II) complex upon precisely controlled clockwise and counterclockwise vortex motions at the air–water interface. Reprinted with permission from [376]. Copyright 2022, Wiley-VCH.
Materials 17 00271 g005
Figure 6. Aggregation behavior of bottlebrush polymers at dynamic interfaces, as formed by electrostatic interaction of carboxylate (poly(acrylic acid)) and ammonium (aminopropylisobutyl polyhedral oligomeric silsesquioxane). Reprinted with permission from [377]. Copyright 2023, American Chemical Society.
Figure 6. Aggregation behavior of bottlebrush polymers at dynamic interfaces, as formed by electrostatic interaction of carboxylate (poly(acrylic acid)) and ammonium (aminopropylisobutyl polyhedral oligomeric silsesquioxane). Reprinted with permission from [377]. Copyright 2023, American Chemical Society.
Materials 17 00271 g006
Figure 7. Dynamic covalent bonding controlled through modulating the droplet shape at the dynamic interface, where anisotropic compartmentalization of the liquid–liquid interface thus occurs, and the pH dependence of the Schiff base reaction leads to reversible toggling from jamming to unjamming in the interfacial assembly. Reprinted with permission from [381]. Copyright 2022, American Chemical Society.
Figure 7. Dynamic covalent bonding controlled through modulating the droplet shape at the dynamic interface, where anisotropic compartmentalization of the liquid–liquid interface thus occurs, and the pH dependence of the Schiff base reaction leads to reversible toggling from jamming to unjamming in the interfacial assembly. Reprinted with permission from [381]. Copyright 2022, American Chemical Society.
Materials 17 00271 g007
Figure 8. Covalent two-dimensional polymer under ambient temperature conditions at the air–water interface with the thickness of the sheet corresponding to that for its monolayer. Reprinted with permission from [382]. Copyright 2016, Wiley-VCH.
Figure 8. Covalent two-dimensional polymer under ambient temperature conditions at the air–water interface with the thickness of the sheet corresponding to that for its monolayer. Reprinted with permission from [382]. Copyright 2016, Wiley-VCH.
Materials 17 00271 g008
Figure 9. Effect of the type of solvent (methanol or N,N-dimethylformamide) in the ligand diffusion solution to control the formation of MOF nanosheets synthesized through spreading the ligand 2,3,6,7,10,11-hexaiminotriphenylene on an aqueous solution containing Ni2+. Reprinted with permission from [388]. Copyright 2023, Chemical Society of Japan.
Figure 9. Effect of the type of solvent (methanol or N,N-dimethylformamide) in the ligand diffusion solution to control the formation of MOF nanosheets synthesized through spreading the ligand 2,3,6,7,10,11-hexaiminotriphenylene on an aqueous solution containing Ni2+. Reprinted with permission from [388]. Copyright 2023, Chemical Society of Japan.
Materials 17 00271 g009
Figure 10. Fabrication of mesoporous thin films with both uniformity and many tens of nanometer pores through a simple method in which a vortex flow is created in a beaker filled with water and carbon nanorings that float on the water surface and are then transferred to a substrate. The resulting thin films were carbonized under an inert gas atmosphere to synthesize carbon nanosheets. Reprinted with permission from [392]. Copyright 2018, Wiley-VCH.
Figure 10. Fabrication of mesoporous thin films with both uniformity and many tens of nanometer pores through a simple method in which a vortex flow is created in a beaker filled with water and carbon nanorings that float on the water surface and are then transferred to a substrate. The resulting thin films were carbonized under an inert gas atmosphere to synthesize carbon nanosheets. Reprinted with permission from [392]. Copyright 2018, Wiley-VCH.
Materials 17 00271 g010
Figure 11. Dynamic liquid crystal transition using interfaces designed to spontaneously delay bulk crystallization and release residual strain at the interface in situ during annealing, spontaneously healing the interface via a dynamic transition to give better perovskite materials. Reprinted with permission from [393]. Copyright 2022, Wiley-VCH.
Figure 11. Dynamic liquid crystal transition using interfaces designed to spontaneously delay bulk crystallization and release residual strain at the interface in situ during annealing, spontaneously healing the interface via a dynamic transition to give better perovskite materials. Reprinted with permission from [393]. Copyright 2022, Wiley-VCH.
Materials 17 00271 g011
Figure 12. Self-adaptive polydimethylsiloxane/TiO2−x coating that can dynamically adapt to volume changes and inhibit dendrite growth with high dynamic adaptability due to the micro-crosslinking of the B-O bonds. Reprinted with permission from [406]. Copyright 2022, Wiley-VCH.
Figure 12. Self-adaptive polydimethylsiloxane/TiO2−x coating that can dynamically adapt to volume changes and inhibit dendrite growth with high dynamic adaptability due to the micro-crosslinking of the B-O bonds. Reprinted with permission from [406]. Copyright 2022, Wiley-VCH.
Materials 17 00271 g012
Figure 13. The kinetically controlled liquid–liquid interfacial precipitation (KC-LLIP) strategy, where ethylenediamine was selected as a covalent cross-linker of C60 and the formation of C60-ethylenediamine products was controlled by the addition of isopropyl alcohol. To introduce the hollow structure, ethylenediamine-sulfur was used as an in situ generated droplet to form the yoke–shell structure, giving porous spheres, string hollow spheres, hollow spheres, and aperture hollow spheres. Reprinted with permission from [418]. Copyright 2022, Wiley-VCH.
Figure 13. The kinetically controlled liquid–liquid interfacial precipitation (KC-LLIP) strategy, where ethylenediamine was selected as a covalent cross-linker of C60 and the formation of C60-ethylenediamine products was controlled by the addition of isopropyl alcohol. To introduce the hollow structure, ethylenediamine-sulfur was used as an in situ generated droplet to form the yoke–shell structure, giving porous spheres, string hollow spheres, hollow spheres, and aperture hollow spheres. Reprinted with permission from [418]. Copyright 2022, Wiley-VCH.
Materials 17 00271 g013
Figure 14. The in situ reaction method to the self-assembly process of C60 molecules and mela-mine/ethylenediamine components in solution, producing a new fullerene assembly, a micron-sized, two-dimensional, amorphous-shaped, regular object, a fullerene rosette (ad). Reproduced under terms of the CC-BY license [420]. Copyright 2022, MDPI.
Figure 14. The in situ reaction method to the self-assembly process of C60 molecules and mela-mine/ethylenediamine components in solution, producing a new fullerene assembly, a micron-sized, two-dimensional, amorphous-shaped, regular object, a fullerene rosette (ad). Reproduced under terms of the CC-BY license [420]. Copyright 2022, MDPI.
Materials 17 00271 g014
Figure 15. Cone–husk fullerene C60 crystals prepared by dynamic liquid–liquid interface precipitation under ambient temperature (ae) and pressure with high sensitivity to acetic acid in quartz crystal resonator sensor electrodes modified with cone–husk fullerene C60 crystals (bottom). Reproduced under terms of the CC-BY license [421]. Copyright 2022, MDPI.
Figure 15. Cone–husk fullerene C60 crystals prepared by dynamic liquid–liquid interface precipitation under ambient temperature (ae) and pressure with high sensitivity to acetic acid in quartz crystal resonator sensor electrodes modified with cone–husk fullerene C60 crystals (bottom). Reproduced under terms of the CC-BY license [421]. Copyright 2022, MDPI.
Materials 17 00271 g015
Figure 16. Supramolecular assembly of fullerene C60 into high aspect ratio quasi-two-dimensional microbelts at room temperature at the dynamic liquid–liquid interface of a carbon disulfide solution of fullerene C60 and isopropyl alcohol. The formed fullerene microbelt can be converted to a mesoporous carbon microbelt with amorphous or graphite backbone structure through carbonizing at 900 °C or 2000 °C, respectively. Reprinted with permission from [422]. Copyright 2017, American Chemical Society.
Figure 16. Supramolecular assembly of fullerene C60 into high aspect ratio quasi-two-dimensional microbelts at room temperature at the dynamic liquid–liquid interface of a carbon disulfide solution of fullerene C60 and isopropyl alcohol. The formed fullerene microbelt can be converted to a mesoporous carbon microbelt with amorphous or graphite backbone structure through carbonizing at 900 °C or 2000 °C, respectively. Reprinted with permission from [422]. Copyright 2017, American Chemical Society.
Materials 17 00271 g016
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

Ariga, K. Materials Nanoarchitectonics at Dynamic Interfaces: Structure Formation and Functional Manipulation. Materials 2024, 17, 271. https://doi.org/10.3390/ma17010271

AMA Style

Ariga K. Materials Nanoarchitectonics at Dynamic Interfaces: Structure Formation and Functional Manipulation. Materials. 2024; 17(1):271. https://doi.org/10.3390/ma17010271

Chicago/Turabian Style

Ariga, Katsuhiko. 2024. "Materials Nanoarchitectonics at Dynamic Interfaces: Structure Formation and Functional Manipulation" Materials 17, no. 1: 271. https://doi.org/10.3390/ma17010271

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