BN Nanostructures and BN-Based Nanohybrids: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor: , , , , ,

Due to its unique physical, chemical, and mechanical properties, such as a low specific density, large specific surface area, excellent thermal stability, oxidation resistance, low friction, good dispersion stability, enhanced adsorbing capacity, large interlayer shear force, and wide bandgap, hexagonal boron nitride (h-BN) nanostructures are of great interest in many fields. These include, but are not limited to, (i) heterogeneous catalysts, (ii) promising nanocarriers for targeted drug delivery to tumor cells and nanoparticles containing therapeutic agents to fight bacterial and fungal infections, (iii) reinforcing phases in metal, ceramics, and polymer matrix composites, (iv) additives to liquid lubricants, (v) substrates for surface enhanced Raman spectroscopy, (vi) agents for boron neutron capture therapy, (vii) water purifiers, (viii) gas and biological sensors, and (ix) quantum dots, single photon emitters, and heterostructures for electronic, plasmonic, optical, optoelectronic, semiconductor, and magnetic devices. BNNPs and BN-based nanohybrids exhibit antibacterial and antifungal activity.

  • hexagonal BN
  • nanostructures
  • nanohybrids
  • fabrication
  • application

1. Introduction

In recent years, the number of works devoted to hexagonal boron nitride (h-BN) nanostructures has grown rapidly, and the scope of their application has expanded significantly. Statistics of publications and citations when searching for keywords “BN nanostructures” in the Web of Science database are shown in Figure 1. The number of annual articles has already exceeded 900, and the number of citations is more than 4000. The main goal of this entry is to provide a critical analysis of the state-of-the-art in the field of h-BN nanostructures based on a review of the most recent works in order to demonstrate their promise in many critical areas of modern science and technology.
Figure 1. Statistics of publications and citations when searching for keywords “BN nanostructures” in the Web of Science database.
The main structural blocks of the sp2-bonded BN are six-membered rings formed by B and N atoms. Regardless of the vertical stacking sequence, the interplanar spacing is ∼0.333 nm. In the simplest AA stacking, each B or N atom sits on top of its B or N counterpart from one layer to another (P6m¯¯¯¯¯2 symmetry). In the case of h-BN with P63/mmc symmetry, every other layer is rotated by 180° around the [0001] axis and every B atom in the parallel hexagons sits on top of a N atom (A-A′ stacking). Rhombohedral BN (r-BN) has three-layer ABC stacking (R3m symmetry), in which, hexagons are displaced by a vector of a/3 in a <011¯0> direction, keeping the same interplanar spacing as h-BN. The sp2-bonded BN is often highly defective, which is a great advantage for some applications (e.g., catalysis, photocatalysis, water splitting and purification, sensor devices) and a strong disadvantage for others (typically optics and electronics). A mixture of h-BN and r-BN crystallites obeying specific orientation relationships [1] is often observed in the sp2-bonded materials. This leads to a local bending of BN atomic planes (this type of BN is called turbostratic BN (t-BN)). Edge dislocations are common within the t-BN. Inside the turbostratic structure, the interlayer distance is increased, which leads to a weakening of the optical and electronic interactions between adjacent layers. Less common is AB of symmetry (so-called Bernal stacking), in which, the centers of hexagons in one layer are vertically aligned with atoms in adjacent layers [2]. AB stacking was observed in chemically exfoliated [3] and synthesized [4] few-layer h-BN.
Hexagonal BN can be prepared as 0D (fullerenes, quantum dots (QDs)), 1D (nanotubes (NTs), nanorods, atomically flat grain boundaries), 2D (monolayer, nanosheets (NSs)), and 3D materials (films, nanoparticles (NPs)). In the 1D configuration, h-BN is of great interest for electronic devices as an atomically sharp AA′/AB stacking boundary [5]. Although graphite-like h-BN is the structural analog of graphene (Gr) and their physical and chemical properties are often compared, many properties, such as their color, electric conductivity, and oxidation resistance, are different. Hexagonal BN is a wide-band-gap (~6.0 eV) insulator [6]. The versatile use of h-BN nanostructures is associated with an interesting combination of properties: a low specific density, high specific surface area, high thermal conductivity and chemical stability, insulation, superior oxidation resistance, a natural optical hyperbolic behavior, and an unusually bright deep-UV emission. In contrast to Gr, which is brittle, single-crystal monolayer h-BN demonstrated high fracture toughness with an effective energy release rate of up to one order of magnitude higher than that predicted by Griffith low and reported for Gr [7]. Recent results have shown that atomically thin h-BN exhibits wetting translucency similar to that of Gr [8]. The piezoelectric characteristics were theoretically predicted in the monolayer h-BN, which has no center of symmetry [9], and subsequently confirmed experimentally in the h-BN monolayer [10] and defective h-BN NTs [11].
Hexagonal BN nanostructures are widely used as additive, thin film, support, encapsulant, carrier, insulator, and barrier material. For nanophotonic, optoelectronic, and sensing applications, it is also considered as a host material for active atomic vacancies/defects. Hexagonal BN is commonly used as a low-loss “dielectric spacer” in many photoelectronic devices.
Currently, h-BN nanomaterials are used in many areas, such as catalysis, photocatalysis, biomedicine (agent for antibacterial, antifungal, antitumor, and boron neutron capture therapy), environmental industry (biodetectors, membranes, and filters), tribology (additive to solid and liquid lubricants), energy (components of cathodes and capacitors in batteries), nanoelectronics, quantum optics and photonics, and deep UV optoelectronics (Figure 2). They are also widely utilized as a reinforcing and/or heat transfer phase in metal, ceramic, and polymer matrix composites, as a modifier of textile materials and soft magnetic composites, and as a filler for heat-insulating aerogels and ionogels. Many of the unique properties of h-BN are associated with point defects, formed either as a result of material production or by subsequent mechanical, chemical, or irradiation treatments. Therefore, their precise controlling is of a great importance. Defect engineering by hydrogen plasma treatment and high-temperature annealing was shown to be an effective tool to control point vacancies and oxygen-related defects [12]. However, the identification of point defects in h-BN is still a challenge. Recently, significant progress has been made in studying the distribution of electric fields at the atomic level using advanced differential phase contrast imaging. Using this method, it is possible to map and measure enhanced electric fields around B monovacancies with respect to an ideal lattice [13].
Figure 2. Applications of h-BN nanostructures.

2. Fabrication and Surface Functionalization

BN flakes with sp2-hybridization were obtained as far back as in 1842 [14]. Hexagonal BNNTs were first synthesized in 1995 using an arc-discharge method [15]. However, the wide application of BNNTs is limited by the low material purity and the inability to control the aspect ratio. A successful step towards obtaining well-dispersed individualized BNNTs is the use of conjugated polymers capable of sorting BNNTs into individual species [16]. In 2016, oxygen containing h-BN nanostructures of various morphologies having hollow and solid cores and smooth and petalled surfaces (with numerous thin-walled h-BN petals) were synthesized using a boron-oxide-assisted chemical vapor deposition (BO-CVD) process [17]. BN nanostructures are obtained by bottom-up and top-down approaches. Several aspects of BN synthesis have been highlighted in recent reviews, mostly in relation to specific applications [18][19][20][21][22][23][24][25].

2.1. Bottom-Up Approach

Disc-shaped h-BNNPs approximately 20 nm in size were synthesized by the microwave heating of a mixture of boric acid and melamine [26]. The use of microwave energy as an alternative heating source made it possible to reduce the reaction time by a factor of three compared to conventional heating. The temperature range for the synthesis reaction was 1130–1210 °C, and the duration of the synthesis was 90 min.
Boric acid and ammonia were used as boron and nitrogen sources in most h-BN synthesis protocols. The fundamental difference in the recently proposed new method is the treatment of ammonia gas with boric acid at room temperature [27]. During the reaction, part of the hydroxyl groups is replaced by amino groups, which leads to the formation of an ammonium borate hydrate compound (NH4)2B4O7 × 4H2O (ABH) containing B-N bonds already at room temperature. The further heating of ABH in a stream of ammonia causes subsequent ammonolysis and dehydration, so this method is called an ammonothermic dehydration process. Synthesis at 550 °C for 24 h made it possible to obtain h-BN nanocrystals approximately 10 nm (Figure 3) in size at a 10% yield. With an increase in the synthesis temperature to 1000 °C, the yield gradually increased. This method paves the way for a new cost-effective and scalable bottom-up approach for the fabrication of h-BN nanocrystals.
Figure 3. Homogeneous BNNSs with an average size of 10 nm synthesized by one-stage low-temperature ammonolysis of boric acid (ammonothermal dehydration method).
Multiwall BNNTs were obtained by CVD at a relatively low temperature of 1050 °C using colemanite as a B precursor and an Fe2O3 catalyst [28]. BN nanoribbons were prepared from BNNTs by applying high pressure [29]. The rupture of NTs led to the formation of new morphological types of h-BN, including nanoribbons encapsulated inside BNNTs. Wrinkled h-BN nanofoam was obtained from graphite by a simultaneous carbothermic reduction and nitriding of boron oxide powder [30]. Interestingly, in this approach, the graphite NSs acted as a carbon source for the reduction reaction and as a rough template for h-BN growth.
The highly crystalline h-BNNPs (FWHM E2g of 11.07 cm−1) with a well-defined nanosheet morphology and large crystal size (2.7–8.4 μm) were successfully synthesized using a polymer-derived ceramics process [31]. The addition of melting point reduction agents (BaF2 and Li3N) made it possible to obtain h-BN powder at a relatively low temperature (1200 °C) and atmospheric pressure. Hexagonal BNNSs with large lateral sizes (a few microns) were also obtained by salt-assisted synthesis [32].
Few-layer h-BN films with micron-sized grains were epitaxially grown by CVD on the surface of a Cu film at a temperature of 1100 °C [33]. The thermal CVD process was carried out in a low-pressure chamber (500 mbar) in a gaseous mixture of NH3 and B2H6 with the addition of H2 and Ar. According to TEM analysis, approximately 92% of h-BN films had one to three atomic layers.
High-quality multilayer h-BN with a controlled thickness (5 to 50 nm) was obtained by the vapor–liquid–solid method using a molten Fe82B18 alloy and N2 as reagents [34]. A thin Fe-B alloy plate was placed over the sapphire and annealed at a temperature above the melting point of the alloy (1250 °C) for 60 min in an Ar/H2 flow (300/50 cm3/min). Further treatment at this temperature was carried out in a N2 flow (300 cm3/min) for 60 min. Liquid Fe82B18 not only supplies boron, but also continuously dissociates nitrogen atoms, supporting the growth of h-BN. The 2D growth was strictly limited by the interface between the liquid Fe82B18 and the sapphire substrate. After cooling, the Fe-B alloy could be easily separated, leaving the h-BN multilayer on the sapphire substrate.
Large single-crystal h-BN with a transverse size of several millimeters is of interest for transistors and photoelectronic devices. Such transparent crystals were obtained by polycondensation, stabilization, and sintering with the addition of preceramic powder [35]. High-quality h-BN crystals were grown at an atmospheric pressure using pure iron as a flux [36]. The narrow vibration Raman E2g peak (7.6 cm−1) and strong phonon-assisted peaks in the photoluminescence spectra evidenced the high quality of the resulting nanomaterial.
The direct growth of high-quality h-BN on dielectric/insulating/transparent substrates is important for electronic and optoelectronic applications. To address this important issue, Bansal et al. [37] studied the growth of h-BN on a (0001) C-plane and (112¯0) A-plane of sapphire (α-Al2O3). The obtained results show that the A-plane is not modified during CVD and, therefore, is a suitable substrate orientation for the growth of h-BN with ABAB staking. Few-layer h-BN was deposited on the sapphire substrate using the ion beam sputtering of an h-BN target in a NH3 atmosphere at a relatively low temperature of 700 °C [38]. Granular nanocrystalline h-BN films were grown on silicon (100) and C-sapphire substrates by low-pressure CVD at 1100 °C using ammonia borane [39]. Hexagonal BN is often used as a substrate for in situ Gr growth to fabricate Gr/h-BN heterostructures for nanoelectronics. Highly ordered epitaxial h-BN was obtained on Gr using a migration-enhanced metalorganic vapor phase epitaxy process [40]. This subject was recently reviewed [41].

2.2. Top-Down Approach

For the BN production on a laboratory scale, physical, chemical, and mechanical methods are used. Mechanical exfoliation in various mediums is the widest group of methods used to obtain BNNSs and BNNPs. Since the interlayer bond force between hexagonal layers is relatively weak, these layers are easy to separate with an external force. The mechanical exfoliation methods include blending, microfluidization, ball milling, and sonication.

2.2.1. Microfluidization

Atomically thin BNNSs with a high aspect ratio of ~1500 were fabricated by microfluidization in a water/ethanol (1:1) solution for 50 cycles at each pressure of 50, 75, 100, and 125 MPa [42]. The exfoliation process takes only 30 min, provides a high yield of 70–76%, and produces approximately 400 mg of low-defect BNNSs. The as-prepared BNNSs were utilized in polymer matrix composites with high thermal conductivity.

2.2.2. Ball Milling

The submicron h-BN powder preliminary synthesized by the high-temperature solid-phase method was exfoliated in a ball mill with the addition of water in a reciprocating mode [43]. This approach demonstrated BNNSs with a thickness of approximately 4–10 nm, a uniform size distribution, and a high yield of 73%. Various saccharides, such as fructose, glucose, maltose, and lactose, were used to improve the efficiency of BN exfoliation during ball milling [44]. Among the studied saccharides, the highest yield of few-layered BNNSs (29.4%) was achieved using glucose. In this mechanochemical exfoliation method, saccharides enhance the applied force of ball milling to promote BN delamination and prevent the reattachment of the nanosheets through their surface functionalization.
When developing ball-milling methods, special attention is paid to studying the influence of the processing medium on the exfoliation process. For example, processing in viscous hydroxyethyl cellulose leads to the formation of BNNSs with a lateral size of 400 nm and a thickness of approximately 2 nm and inhibits NPs agglomeration [45]. Yusupov et al. [46] obtained a high yield of uniform and undeformed BNNSs with an average size of 300 × 600 nm2 and a thickness of approximately 20–50 nm by the ball milling of micron-size BN particles in ethylene glycol with a high density of 1.1 g/m3. Importantly, the milling balls move along the periphery of the chamber, in contact with the walls. This regime creates tangential forces acting on BN particles and leads to their exfoliation. Ball milling with the addition of 2-furoic acid made it possible to achieve a very high yield (~98%) of BNNSs with a thickness of ~2 nm and a lateral size of up to 2 μm [47]. Ultrathin (with average thickness of 3.5 nm) h-BNNSs noncovalently functionalized by ionic liquid were prepared using a liquid ball milling method [48].

3. Catalysts

3.1. Heterogeneous and Homogeneous Catalysts

Many heterogeneous catalysts contain an active phase, a carrier, and promoters. The influence of the support on the catalytic activity, selectivity, and stability of heterogeneous catalysts can be both indirect and direct. The substrate provides not only a high surface area for the active NPs (Figure 4) but also, due to the metal–support interaction, prevents the NPs from agglomeration during temperature-activated catalytic reactions. Since active NPs are usually less than 10 nm in size, a chemical interaction between the carrier surface and the atoms of the active phase often becomes noticeable. This leads to a change in the electron density distribution of the active phase (this effect is called a strong metal–support interaction (SMSI)). Finally, the support material has a direct effect on the course of catalytic reactions due to the interaction of the reaction components with the support. The h-BN active edges (shown by arrows in Figure 4c) can intensify the chemical interaction with the reagents. Thus, the carrier surface provides new active sites for the activation of molecules.
Figure 4. Heterogeneous FePt/h-BN (a,b) and Ag/h-BN (c) catalysts. Arrows show active h-BN edges.
Although metal oxides are the most studied class of materials for many catalytic reactions, in recent years, h-BN has received increased attention as a support of catalytically active species. BN nanomaterials are attractive due to their large specific surface area, high resistance to sedimentation in liquid media, relatively high thermal stability and corrosion resistance, and the ability to tune their structure and properties by doping, creating defects, or by surface functionalization. The use of h-BN as a carrier can increase the activity of the catalyst compared to conventional supports. Due to their layered structure, h-BN provides a high density of attached metal NPs, which leads to increased catalytic activity [49][50]. BN has unshared electron pairs localized on nitrogen atoms, resulting in a polarized state. In addition, the electronic properties of h-BN depend very strongly on impurities, the introduction of which makes it possible to reduce the band gap. One of the paradoxes of h-BN is that, being a highly inert material, it has its own catalytic activity. This was clearly shown in the reaction of the oxidative dehydrogenation of propane [51]. The available experimental and theoretical data indicate that the h-BN surface acts as a driver of conversion [52][53][54].
The combination of different functional properties in one material can lead to additional advantages in the implementation of catalytic processes. For example, by creating core–shell Co@h-BN structures, it was possible to achieve a high activity and structural stability of Co NPs, and their ferromagnetic nature made it possible to carry out the hydrogenation of nitroarenes in aqueous media [55]. Hexagonal BN acted as a protective layer, reducing the probability of Co NP agglomeration and, as a result, contributed to an increase in the active surface and adsorption of reagent molecules. The FeNiCo/h-BN material was characterized as a highly active and selective catalyst for the hydrogenolysis of hydroxymethylfurfural [56]. The high BE between the FeNi and Co NPs and the h-BN support ensured the stability of the catalyst and the possibility of its reuse. The surface chemistry of the support in the heterogeneous catalyst significantly affects the material performance. It was shown that a large number of B-O and N-H groups in the Co/h-BN catalyst leads to an enhanced CoO/h-BN interaction, which prevents the reduction of Co oxide to active metal NPs [57]. The low amount of functional groups leads to the formation of larger Co NPs due to the reduced BE. An increased adsorption of cinnamaldehyde was associated with a high concentration of NH groups. The interaction of h-BN with active NPs can be enhanced by additional surface functionalization. Thus, the stability of catalytically active sites deposited on the h-BN surface during the oxidative desulfurization of the fuel was increased by modifying h-BN with an ionic liquid [58]. The modification also led to an increase in the efficiency of catalyst recirculation due to the enhanced interaction of the acid with the h-BN surface. The excellent catalytic properties of BN-based catalysts are largely due to the presence of defects. It was shown that the B-OH bonds formed at the edges of defective h-BN flakes serve as anchor centers for Cu NPs [59]. The strong interaction force resulted in a high resistance of Cu to sintering, which is usually difficult due to the low Hutting temperature of copper. The synthesized catalysts were tested in the reaction of ethanol dehydrogenation and showed a better ability to selectively form acetaldehyde compared to oxide counterparts (silica and alumina). The observed effect is explained by the difference in acetaldehyde binding: the h-BN-based catalysts were able to adsorb ethanol strongly, whereas the interaction with acetaldehyde was weak. At the same time, oxide-supported catalysts showed a strong interaction with both ethanol and acetaldehyde. This behavior resulted in an excellent selectivity of h-BN-based systems in the production of acetaldehyde.
The high catalytic efficiency of many heterogeneous materials is associated with the small size of active centers. The small size of Ag NPs and their maximum density on the h-BN surface were shown to be key parameters determining the high catalytic activity. Even a slight increase in the Ag NP size led to a noticeable deterioration in the catalytic performance in terms of offset and full conversion temperatures [60]. Pt cluster-supported BNNSs (with 100 ppm of Pt) were characterized as a highly efficient catalyst for propane dehydrogenation with high propane conversion (~15%) and propylene selectivity (>99%) at a relatively low reaction temperature (520 °C) [61]. To reduce the size of Pd NPs on an h-BN support, it was proposed to use an intermediate MgO layer [62]. It was found that, in addition to reducing the Pd size, Mg2+ ions partially penetrate into the h-BN crystal lattice, filling B vacancies. This contributed to an increase in oxygen adsorption and had a positive effect on the catalytic activity. The advantage of the metal oxide/BN interface was also demonstrated for the Ni/CeO2/h-BN system [63], in which, Ni NPs were imbedded between cerium oxide and BN. The stability of Ni NPs was increased due to the metal/support interaction, which permitted reducing coke formation in the dry methane reforming reaction. The high CO2 conversion rate observed on the Au/BN and Pt/BN catalysts is associated with better CO2 adsorption on the oxidized BN surface. A charge density distribution at the Pt/h-BN interface increases oxygen absorption, thereby accelerating oxygen-associated chemical reactions [64].
Fe3O4/BN, Fe3O4(Pt)/BN, and FePt/BN nanohybrids were obtained via polyol synthesis in ethylene glycol [65]. BN supporting bimetallic FePt NPs demonstrated a significantly higher CO2 conversion rate compared to Fe3O4/BN and Fe3O4(Pt)/BN counterparts and an almost 100% selectivity to CO, whereas catalysts with Fe3O4 NPs showed a better selectivity to hydrocarbons. An important result is the formation of core–shell h-BN@FePt structures upon heating, which prevents the agglomeration of catalytically active NPs.
In the case of SMSI, the metal NP can be encapsulated in a carrier material. This can have both a positive and negative effect on the catalyst performance. Hexagonal BN rarely exhibits the SMSI effect due to its relative inertness; however, BO bonds present as surface functional groups can act as SMSI sites. SMSI between FeOx and BOx was observed in Fe/h-BN catalysts, which led to the coordination of reduction and oxidation processes during the oxidative dehydrogenation of ethylbenzene to styrene [66]. As with oxide supports, SMSI can be customized with redox cycles in Pt/h-BN catalysts. It has been shown that the oxidative treatment of Pt/h-BN at temperatures above 520 °C leads to the formation of BOx species and their strong interaction with metal NPs [67]. The formation of thin layers of boron oxide on top of Pt NPs was observed, which blocked low-coordinated Pt centers. Encapsulated Pt NPs were resistant to sintering, and the particle size remained unchanged during catalysis. The Pt/h-BN catalysts showed better stability in the propane dehydrogenation reaction compared to the Pt/Al2O3 counterpart due to the lower rate of coke formation. The surface oxygen-terminated groups of h-BN can lead to too strong an interaction between the active phase and the carrier and adversely affect the reducibility of metal oxide NPs. To overcome this drawback, Fe/h-BN catalysts doped with Cu and Mn were synthesized for the Fischer–Tropsch process [68]. Doping Fe/h-BN with Mn promotes the reduction and formation of active iron carbides, while Cu lowers the Fe-O reduction temperature, creating more active sites for H2 dissociation. The observed synergistic effect of the promoters increased the catalyst activities without reducing their stability. However, in some cases, too strong an active phase/h-BN interaction is unfavorable, as in the case of an acid–base catalytic reaction such as the synthesis of dimethyl ether by dehydration of methanol [69]. This reaction usually proceeds on Brønsted acid catalysts and is therefore sensitive to the number and strength of the acid sites. In this case, h-BN is not an ideal carrier among SiO2, TiO2, ZrO2, Al2O3, and CeO2 due to the strong interaction between tungstosilicic acid and the support, which leads to a decrease in proton mobility and catalyst efficiency.
Elemental doping is an effective approach to band gap engineering. Doping with carbon makes it possible to control the chemical activity of h-BN by changing the electron density near the Fermi level. Carbon-doped h-BN showed increased activity in the condensation of benzaldehyde with malononitrile to benzylidenemalononitrile, surpassing the undoped h-BN and C3N4 counterparts [70]. The high activity is explained by the decrease in the desorption barrier of the reaction products due to carbon doping. The addition of Se to h-BN has been shown to narrow the band gap and improve carrier generation and separation [71].
A number of studies have shown that the direct interaction of the h-BN support with the reaction products affects the catalytic characteristics. A comparison of Pd NP catalysts supported on h-BN, Al2O3, and MgO showed that the adsorption of maleic anhydride on Pd centers is improved as a result of the Pd to carrier interaction, and the Pd/h-BN system is characterized by increased adsorption on h-BN, which has a direct effect on the increase in catalytic activity [72]. The formation rate of succinic acid 6000 g/gkat/h at a selectivity of 99.7% on the Pd/h-BN catalyst was higher than on their oxide-based counterparts. The catalytic activity can be increased due to the formation of hydroxyl groups on the h-BN surface [73]. For example, an increase in the sonication time made it possible not only to increase the loading of metal, which led to an increase in the reduction rate of 2-nitroaniline, but also to increase the activity of catalysts due to the formation of hydroxyl surface groups [74]. However, in some cases, the relatively inert h-BN surface provides high selectivity for target products. A comparison of Pt-catalysts on various supports during the hydrogenation of cinnamic aldehyde to cinnamic alcohol showed that the absence of acid–base sites on the h-BN surface makes it an ideal candidate as a support for this reaction, since the process proceeds through a simple non-dissociative adsorption of cinnamic aldehyde. This results in a selectivity of over 85% towards cinnamyl alcohol. In the case of Al2O3 and SiO2 supports with a large number of acid–base sites, the selectivity is hindered due to the multiple-adsorption regime of the reagent.
Hexagonal BN can affect the electronic structure of deposited metal NPs. It was shown that the location of Ru on the h-BN edges leads to an increase in the hydrogenation activity due to the enhancement of interfacial electronic effects between Ru and the BN surface [75]. Electron-enriched NPs showed high activity in the production of primary amines from carbonyl compounds without the need for excess ammonia. For practical application, it is very important to control the electron density. An important step in this direction is the work of Zhu et al. [76], showing that the change in the electron density of Pt NPs is possible by adjusting the B and N vacancies in h-BN. Pt NPs are located at the site of boron and nitrogen vacancies acting as Lewis acids and Lewis bases, respectively. Thus, the synthesis of a Pt/h-BN nanohybrid with a predominance of nitrogen vacancies leads to an interfacial electronic effect that promotes O2 adsorption, reduces CO poisoning, and increases the overall activity and stability of the catalyst in the CO oxidation reaction.
The 2D morphology of h-BN allows it to be successfully utilized as a homogeneous catalyst in liquid heterogeneous catalytic reactions due to its good dispersibility and, hence, high specific surface area. To obtain a metal-free catalyst for the oxidative desulfurization of aromatic compounds, N-hydroxyphthalimide (NHPI) was covalently grafted to the h-BN surface [77]. The strong interaction between NHPI and h-BN results in a high catalyst activity, selectivity and recyclability. Rana et al. [78] covalently grafted a copper complex onto the surface of exfoliated BNNSs. To ensure the stability of the complex, 3-aminopropyltriethoxysilane was used as a covalent linker. Leaching tests showed that there was no depletion of Cu from the catalyst surface. The materials showed exceptional activity in the azide-nitrile cycloaddition reaction to produce the pharmaceutically important 5-substituted 1H-tetrazole.
Single-atom catalysts (SACs) that combine the advantages of homogeneous and heterogeneous catalysts are of great interest due to exceptionally high catalytic activity in a wide variety of industrially important catalytic reactions [79]. To commemorate the 10th anniversary of the introduction of the term SACs, a recent review examined various single-atom–host combinations and related applications [80]. From an economic point of view, it is highly desirable to reduce the content of expensive metals of the Pt group without compromising activity, selectivity, and stability. The chemical activity of h-BN can be fine-tuned using lattice defect engineering. For example, during the cryogenic grinding of h-BN powders, vacancies are formed that can serve as active centers for the spontaneous reduction of metal cations [81]. The authors showed the possibility of the formation of single atoms and clusters of Ag, Au, Pt, Cu, and Fe on h-BN defects. The nitrogen-containing B vacancy in h-BN can effectively anchor and confine Pd atoms [82]. Hexagonal BN was shown to be a suitable material for dispersing Cu-Pt clusters due to the abundant B-O species on its surface [61]. Nanoscale clusters (1–2 nm) were active and stable in the catalytic dehydrogenation of propane to propylene.
It is important to note that the choice of support material should be carried out taking into account the catalytic reaction mechanism. For example, the study of catalysts in the oxidative dehydrogenation of alcohols showed that the support material should be able to accept and conduct electrons between catalytically active NPs [83]. From this point of view, h-BN was inferior to C and TiO2 due to its high band gap.
Hexagonal BN also finds its application in a relatively new catalytic direction: microwave catalysis. Mo2C/h-BN nanomaterials were proposed as highly active H2S decomposition catalysts [84]. However, the role of the support in this area of catalytic processes is still poorly understood.

3.2. Photocatalysts and Electrocatalysts

Photocatalysis and electrocatalysis are alternative processes for efficient chemical transformations. In contrast to thermal catalysis, additional energy enters the system from an external source, which makes it possible to overcome the thermodynamic barrier and shift the chemical equilibrium towards the reaction products. In photocatalytic and electrocatalytic processes, materials, in addition to the ability to adsorb and activate chemical reagents, must be able to conduct electrons and be activated under the action of external energy. From this point of view, h-BN is not an ideal material. Due to the high band gap, a relatively high energy is required to excite an electron and transfer it from the valence band to the conduction band. However, the combination of h-BN with other materials proved to be effective in achieving high activity and stability.
The efficiency of the photodegradation of organic pollutants under visible light irradiation can be increased in the presence of h-BN. For example, the addition of h-BN to the ZnFe2O4 photocatalyst decreased the recombination rate of electron–hole pairs, which affected the efficiency of Congo red and tetracycline degradation [85]. The incorporation of h-BN into 2D Bi2WO6 flakes increased the rate of charge separation and reduced the electron–hole recombination [86]. This increased the ability to degrade antibiotics by 1.87 times. The 2D/2D h-BN/Bi2WO6 heterostructures also showed high activity in the antibiotic degradation [87]. MoS2/h-BN/reduced graphene oxide (rGO) composites were developed and tested for water splitting under sunlight [88]. The simultaneous presence of BN/rGO, which has a high structural stability and a large number of surface centers, and MoS2 with a small band gap, provides a synergistic catalytic effect. The hydrogen evolution rate of 1490.3 µmol/h/g was 58.2 and 12.2 times higher than that of the BN/rGO and MoS2 counterparts. The high catalytic activity and selectivity of h-BN-based photocatalysts in the reaction of nitrate reduction under UV radiation was explained by the formation of photoelectrons and CO2 radicals [89]. BN provided a high concentration of adsorption sites with B atoms acting as Lewis acids, and its electronic configuration was well-suited for efficient nitrate reduction. The hexagonal BN support was shown to provide a sink of excited electrons from the surface of catalytically active Cu3P nanoparticles, making them active on the h-BN surface for the adsorption of reagent molecules [90]. BN was characterized as a promising catalyst for organic degradation via the activation of peroxymonosulfate. The defect-driven non-radical oxidation of porous h-BN nanorods was proposed as the main mechanism of sulfamethoxazole degradation via the formation of singlet oxygen (1O2) [91]. Used BN was easily regenerated upon heating in air, which completely recovered the B–O bonds.
a-MoSxOy/h-BNxOy nanomaterial with a tuned MoSxOy zonal structure for photoinduced water splitting was developed [92]. The nanohybrid showed high activity in the photocatalytic degradation of methylene blue (MB) (5.51 mmol g−1 h−1 under illumination with a mercury lamp), four times higher than that of known non-metallic catalysts. The obtained photocatalyst is very stable and can be reused. Oxygen-substituted BN has good wettability, and the interaction of oxygen defects with sulfur leads to the formation of S-doped BN, which has a high photodegradation efficiency when illuminated with visible light [93].
To adapt the commercial PtRu/C electrocatalyst as an anode in fuel cells with a proton-exchange membrane, metal NPs on a carbon support were encapsulated in few-layers h-BN [94]. The h-BN shell weakened the CO adsorption on the PtRu surface and increased the resistance to CO in H2-O2 fuel cells. The structural stability of Mo2N electrocatalysts for the nitrogen reduction reaction was improved by combining molybdenum nitride with h-BN [95]. The optimal catalyst exhibited a NH3 yield rate of 58.5 µg/h and a Faraday efficiency of 61.5%. By designing h-BN defects, an optimum conversion rate was achieved at a lower overvoltage. Carbon-doped h-BN was used to create enzyme-like single-atom Co electrocatalysts for the dechlorination reaction [96]. Locally polarized B-N bonds played a key role in the adsorption of organochlorine compounds, improving the electrocatalytic activity. No such effect was observed on carbon and graphite supports doped with nitrogen.
Heterojunctions in h-BN-doped BiFeO3 and MnFeO3 perovskite-based catalysts exhibited an extended visible light range, reduced band gap energy, and low recombination rate, and had excellent photocatalytic activity towards various antibiotics and dyes. The good photocatalytic performance of heterojunctions was explained by an extended excitation wavelength and a slow recombination rate of charge carriers [97].
In conclusion, researchers note the important role of the h-BN carrier, which, interacting with surface NPs, can increase their structural stability and change their electron density, which affects the catalytic activity, selectivity, and stability. Hexagonal BN can also be directly involved in chemical reactions, creating additional sites for component activation. All of these effects can have a positive or negative effect depending on the type of reaction, which is quite consistent with the well-known Sabatier–Balandin principle. Therefore, when designing a catalytic system, one should take into account the mechanisms of catalytic reactions.

4. Materials for Biomedicine and Improvement of Quality of Life

4.1. Biocompatibility and Dose-Dependent Toxicity

The biocompatibility and bio-application of BN nanomaterials have recently been reviewed [98]. Available data indicate that the biocompatibility of BN depends on the concentration, size, and shape of BNNPs. In earlier studies, the cytocompatibility of BN nanostructures was evaluated in relation to various cell cultures, such as kidneys, epithelium, human skin, ovarian, bone tissue, human carcinoma, human osteosarcoma, human lung epithelial adenocarcinoma, human neuroblastoma, and others. The results show that a concentration of BNNPs less than 40 μg/mL is safe for the vast majority of cell lines, regardless of the size (5–200 nm) and shape (nanospheres, NSs, NTs). Moreover, a low concentration of BNNPs can stimulate cell proliferation. The addition of BNNPs to differentiated NT-2 cells increased their viability by 6% (6.25 µg/mL) and 13% (3.12 µg/mL) [99]. Low BN concentrations (10–100 μg/mL) increased the viability of human umbilical vein endothelial cells by 118% (p < 0.05), while a NP concentration of 150 and 200 μg/mL had no significant effect on cell metabolism [89]. This effect can be explained by the influence of NPs at their low concentration on oxidative stress. Radicals (°OH, °OOH) can react with proteins, enzymes, nucleic acids, and other cell biomolecules, causing their damage and cellular apoptosis. Recently, it was shown that h-BN can slowly dissolve in phosphate-buffered saline (PBS) and lysosome mimicking solution, while the released boron can form boric acid [100], which exhibits antioxidant and antiapoptotic effects [101][102][103]. Boric acids and their esters are cleaved by H2O2 and other reactive oxygen species (ROS) to form the corresponding alcohol and boric acid, which are considered nontoxic to human cells [104]. The effect of BNNPs on cell viability (mHippo E-14 cells) was assessed in the presence of doxorubicin (DOX) at a concentration high enough to cause cellular stress but low enough not to kill all cells [105]. The obtained results indicate that h-BN reduces DOX-induced oxidative stress on cells at a concentration of 44 μg/mL.
The dose-dependent toxic effect of h-BN NPs was studied in vivo in albino rats (Wistar) by measuring thiol/disulfide homeostasis, lipid hydroperoxide levels, and myeloperoxidase and catalase activity [106]. After the intravenous administration of various BN doses, hematological and biochemical parameters did not change up to concentrations of 800 µg/kg. However, at doses of h-BNNPs greater than 1600 μg/kg, significant damage to the liver, kidneys, heart, spleen, and pancreas was observed. The results showed that h-BNNPs with a diameter of 120 nm are non-toxic and can be used in biomedical applications at low doses of 50 to 800 µg/kg.

4.2. Antibacterial and Antifungal Activity

BNNPs and BN-based nanohybrids exhibit antibacterial and antifungal activity; some recent results are presented in Table 1.
Table 1. Antibacterial activity of BN NPs and BN-based nanohybrids.
Material BNNP Content (%) Pathogens Ref.
PNMPy-BNNPs 10.0 E. coliS. aureusP. aeruginosaE. faecalis [107]
LDPE-BNNPs 5.0–20.0 E. coliS. aureusP. aeruginosa,
S. epidermidis
[108]
PHA/CH-BNNPs 0.1–1.0 E. coli K1
Methicillin-resistant S. aureus
[109]
QAC-BNNPs-PP 3.0–10.0 E. coli Carolina #155065A
S. aureus Carolina #155556
[110]
CEL-BNNPs 1.0–3.0 E. coli K12 (ATCC 29425)
S. epidermidis ATCC 49461
[111]
  MIC of BN (mg/mL)    
BNNPs 15 Multidrug resistant E. coli (12 strains) [112]
BNNPs 1.62 S. mutans 3.3 [113]
400 S. mutans ATTC 25175
400 S. pasteuri M3
3.25 Candida sp. M25
BNNPs 256 E. coli [114]
128 B. cereus
128 S. aureus
128 E. hirae
128 P. aeruginosa
256 L. pneumophila subsp. pneumophiia
256 C. albicans
BNNSs 100 E. coli DH5α [115]
Poly(N-methylpyrrole (PNMPy), polyhydroxyalkanoate (PHA), cellulose (CEL), chitosan (CH), polypropylene (PP), quaternary ammonium salt (QAC), low-density polypropylene (LDPE).
The following bacteria and fungi suppression mechanisms are noted: chemical (ROS formation [105], slow BN degradation with the formation of boric acid), structural (contact-killing by sharp BN surface structures [116] or an increase in specific surface area, leading to greater damage to bacteria [117][118]), and electrostatic (negatively charged BNNPs interact with bacterial cell walls). A contact-killing bactericidal effect of BNNPs was compared with the toxic effect of gentamicin-loaded BNNPs at the minimal inhibitory concentration of 150 mg/mL [116]. To enhance antibacterial activity, various hybrid NPs, such as BNNPs-Ag [112], BNNPs-ZnO [119], BNNPs-Cu [120], BNNPs-Zr [121], and antibiotic-loaded BNNPs [112] and films [116], have been developed. The combination of bactericidal ions and antibiotics allows one to achieve a pronounced synergistic effect. For example, gentamicin-loaded BNNPs exhibited high bactericidal activity against S. aureusP. aeruginosa, and 38 types of the E. coli strains, including multidrug-resistant ones [112]. For the rest of the tested E. coli strains, the Ag NPs-containing nanohybrids showed superior bactericidal efficiency. In addition, Ag/BN, amphotericin B/BN, and amphotericin B/Ag/BN nanohybrids revealed high fungicidal activity.
Recently, antibiotic-like activity of h-BNNSs combatting antimicrobial-resistant bacteria through a Z-ring constriction damage mechanism has been reported [122]. BNNSs target key surface proteins (FtsP, EnvC, and TolB) and disrupt Z ring constructions.

4.3. Drug Delivery

BNNPs are promising carriers for targeted drug delivery. Since most therapeutic agents (TA) have a complex structure and/or charge, the formation of h-BN/TA complexes is possible due to van der Waals and/or electrostatic interactions between the components. Hexagonal BN is a good adsorbent for a large number of TA.
The high dispersibility and low reactivity of the BN surface are important factors when introducing NPs into the bloodstream to avoid thrombus formation and hemolytic activity [123]. Since the h-BN surface is hydrophobic and the material has high absorbent properties, surface functionalization by chemical treatment or coating with polymers is necessary. For example, stable BNNTs and h-BN dispersions were obtained using an O3-based advanced oxidation process followed by polyethylenimine functionalization [124].
The cytotoxicity of exfoliated h-BN flakes functionalized with hydroxyl groups (BN-OH) was studied in various models: insect hemocytes (in vivo), human erythrocytes, and mouse fibroblasts (in vitro) [125]. Long-term immunoassays showed that BN-OH, despite the absence of hemocytotoxicity, impaired nodulation, the most important cellular immune response in insects. Hemocytes exposed to BN-OH and then to bacteria differed in morphology and adhesiveness from hemocytes exposed to bacteria alone and exhibited the same morphology and adhesiveness as control hemocytes. Thus, the BN-OH-induced decrease in nodularity may be the result of a decrease in the ability of hemocytes to recognize bacteria, migrate to them, or form microaggregates around them, which can lead to immune system dysfunction when infected by pathogens. Long-term in vivo studies are still needed to unambiguously confirm that h-BN is biocompatible and can be utilized as a drug delivery platform or for bioimaging.
In addition to surface functionalization, which improves dispersibility and biocompatibility, the attachment of specific surface ligands or complexes is necessary to ensure the selectivity of drug delivery carriers. Using the self-assembly approach, a-BN-based hybrid NPs containing DOX and a conjugate of a mixture of folic acid (FA) and chitosan (CH) were prepared [126]. N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride was used as the crosslinking agent. The DOX-loading capacity in a-BN with and without FA-CH was 7.52 ± 0.32% and 6.24 ± 0.58%, respectively. For in vivo cell visualization, DOX was replaced with cyanine 5.5 as a label. Mice injected with a-BN-Cy5.5@FA-CH showed intense fluorescence in the tumor region, indicating the effectiveness of FA-CH conjugates for targeted drug delivery.
For the treatment of the most aggressive type of brain cancer (glioblastoma multiforme), BNNTs were coated with cell membranes extracted from glioblastoma cells [127]. Effective targeted drug delivery is based on the homotypic recognition of tumor cells included in the carrier. The cell membrane-conjugated BN complexes acted specifically and effectively killed cancer cells without affecting healthy brain cells.
BN nanostructures are also being used to create theranostic systems with simultaneous therapeutic and diagnostic capabilities. For example, h-BNNSs were used to create a multifunctional theranostic platform containing DNA oligonucleotide and copper (II) phthalocyanine (Pc) [128]. The nanohybrids showed the effective drugs accumulation in tumor cells and remarkable photodynamic therapy efficiency with minimized damage to normal tissues. The Cu Pc molecule was used as a photosensitizer in photodynamic therapy and as a sensitive and accurate diagnostic probe for the in situ monitoring and visualization of miR-21 by surface-enhanced Raman spectroscopy.

4.4. Boron Neutron Capture Therapy

Natural elemental boron contains approximately 20% stable Boron-10 isotopes sensitive to neutron irradiation and 80% Boron-11 isotopes [129]. Boron neutron capture therapy (BNCT) is an emerging and non-invasive cancer treatment strategy based on the selective accumulation of boron compounds in tumor cells and subsequent exposure to a thermal neutron beam. The fundamentals of BNCT and clinical applications were recently reviewed by Malouff et al. [130]. Note that clinically approved B-containing drugs (boronophenylalanine and mercaptoundecahydrododecaborane) have a low cumulative capacity and high cost [131].
BNNPs are promising boron carriers containing 50% B (B-10 or B-11) for BNCT. The main challenge is to achieve sufficient boron accumulation in tumor cells and ensure the subsequent degradation of nanocarriers in order to avoid in vivo toxicity. To help solve this important problem, BNNPs were coated with phase-transitioned lysozyme, which protects nanocarriers from hydrolysis during circulation in blood and can be readily removed with vitamin C after BNCT [132]. The effectiveness of tumor treatment can be enhanced by a combination of antibiotics and neutron therapy. The tumor size decreased by 48.2 ± 17.0% (DOX/BNNSs), 64.1 ± 10.9% (BNNSs + neutron irradiation) and 94.6 ± 2.1% (DOX/BNNSs + neutron irradiation) [133]. In addition, BN nanostructures can be further doped or functionalized to endow them additional functionality. For example, 64Cu radioisotope was added to BNNTs as a biological marker for diagnostic purposes [134]. Cu as a chelating element can potentially bind to peptides and other important biological molecules, such as antibodies, proteins, and nanoparticles.

4.5. Tissue Engineering

BN nanostructures can be used to strengthen tissue-engineering scaffolds and stimulate cell adhesion and proliferation for efficient tissue regeneration. The addition of BNNTs to tricalcium phosphate scaffolds not only increased mechanical properties but also enhanced bone morphogenetic proteins, the osteogenic differentiation of mesenchymal stem cells, Runx2 expression, and new bone formation [135]. The doping of porous BN nanofibers to poly(vinyl alcohol) (PVA) hydrogels was shown to improve its mechanical properties, swelling ability, and thermal stability [136]. Exfoliated BN was used as filler in 3D porous polylactic acid (PLA) scaffolds. The addition of 0.1% BN to PLA scaffolds increased the polymer degradation temperature from 298 °C to 359 °C and improved the swelling from 80 to 118 [137]. The presence of BN in 3D-printed PLA/BN scaffolds did not affect the viability of MG-63 and MC3T3-E1 cells (MTT assay) after 4 or 7 days. Adhesion, proliferation, and cell mineralization on PLA/BN composites were higher compared to the BN-free counterpart.
PVA/h-BN/bacterial cellulose scaffolds for bone tissue engineering were fabricated using 3D printing [138]. The obtained composites showed 100% human osteoblast cell viability after 72 h incubation. The h-BN was used as a filler due to its superior thermal and mechanical properties. The addition of 10% BNNPs to Poly(N-methylpyrrole) increased the antibacterial effect against four types of strains (Escherichia coliStaphylococcus aureusPseudomonas aeruginosa, and Enterococcus faecalisbacterial) [107]. Kirschner wires were coated with h-BN films by magnetron sputtering and studied in the treatment of fractures in adult male Wistar albino rats [139]. According to microcomputed tomography, BN-coated implants showed better healing characteristics compared to the control group.
A comparison of the physicochemical properties of a bioceramic-based root canal sealer used in endodontics reinforced with various nanomaterials (multi-walled carbon nanotubes (CNTs), titanium carbide, and h-BN) showed that the BN-doped composites have minimal initial (at 1%BN) and final (at 2%BN) setting times [140]. The potential use of BNNPs for wound healing was studied using human umbilical vein endothelial (HUVE) and human dermal fibroblast (HDF) cells and compared with boric acid, which is one of the degradation products of h-BN [100]. The proliferation and migration of HUVE and HDF cells were significantly higher in BN-treated cultures compared to boric-acid-treated cells. Hexagonal BNNPs showed no cytotoxicity at concentrations of 25–200 μg/mL. Composites based on chitosan polyhydroxyalkanoates doped with BNNPs (0.1–1.0%) were obtained using a simple solvent casting technique [109]. Compared to a negative control, the addition of BNNPs reduced the percent viability of E.coli K1 and multidrug-resistant S. aureus strains.

This entry is adapted from the peer-reviewed paper 10.3390/nano12162810

References

  1. Shtansky, D.V.; Tsuda, O.; Ikuhara, Y.; Yoshida, T. Crystallography and Structural Evolution of Cubic Boron Nitride Films during Bias Sputter Deposition. Acta Mater. 2000, 48, 3745–3759.
  2. Sponza, L.; Amara, H.; Attaccalite, C.; Latil, S.; Galvani, T.; Paleari, F.; Wirtz, L.; Ducastelle, F. Direct and Indirect Excitons in Boron Nitride Polymorphs: A Story of Atomic Configuration and Electronic Correlation. Phys. Rev. B 2018, 98, 125206.
  3. Warner, J.H.; Rümmeli, M.H.; Bachmatiuk, A.; Büchner, B. Atomic Resolution Imaging and Topography of Boron Nitride Sheets Produced by Chemical Exfoliation. ACS Nano 2010, 4, 1299–1304.
  4. Sutter, P.; Lahiri, J.; Zahl, P.; Wang, B.; Sutter, E. Scalable Synthesis of Uniform Few-Layer Hexagonal Boron Nitride Dielectric Films. Nano Lett. 2013, 13, 276–281.
  5. Park, H.J.; Cha, J.; Choi, M.; Kim, J.H.; Tay, R.Y.; Teo, E.H.T.; Park, N.; Hong, S.; Lee, Z. One-Dimensional Hexagonal Boron Nitride Conducting Channel. Sci. Adv. 2020, 6, eaay4958.
  6. Cassabois, G.; Valvin, P.; Gil, B. Hexagonal Boron Nitride is an Indirect Bandgap Semiconductor. Nat. Photonics 2016, 10, 262–266.
  7. Yang, Y.; Song, Z.; Lu, G.; Zhang, Q.; Zhang, B.; Ni, B.; Wang, C.; Li, X.; Xie, X.; Gao, H.; et al. Intrinsic Toughening and Stable Crack Propagation in Hexagonal Boron Nitride. Nature 2021, 594, 57–61.
  8. Wagemann, E.; Wang, Y.; Das, S.; Mitra, S.K. On the Wetting Translucency of Hexagonal Boron Nitride. Phys. Chem. Chem. Phys. 2020, 22, 7710–7718.
  9. Duerloo, K.-A.N.; Ong, M.T.; Reed, E.J. Intrinsic Piezoelectricity in Two-Dimensional Materials. J. Phys. Chem. Lett. 2012, 3, 2871–2876.
  10. Ares, P.; Cea, T.; Holwill, M.; Wang, Y.B.; Roldán, R.; Guinea, F.; Andreeva, D.V.; Fumagalli, L.; Novoselov, K.S.; Woods, C.R. Piezoelectricity in Monolayer Hexagonal Boron Nitride. Adv. Mater. 2019, 32, 1905504.
  11. Kundalwal, S.I.; Choyal, V. Enhancing the Piezoelectric Properties of Boron Nitride Nanotubes through Defect Engineering. Phys. E 2021, 125, 114304.
  12. Xiao, Y.; Yu, H.; Wang, H.; Zhu, X.; Chen, L.; Gao, W.; Liu, G.; Yin, H. Defect Engineering of Hexagonal Boron Nitride Nanosheets via Hydrogen Plasma Irradiation. Appl. Surf. Sci. 2022, 593, 153386.
  13. Cretu, O.; Ishizuka, A.; Yanagisawa, K.; Ishizuka, K.; Kimoto, K. Atomic-Scale Electrical Field Mapping of Hexagonal Boron Nitride Defects. ACS Nano 2021, 15, 5316–5321.
  14. Balmain, W.H. Bemerkungen über die Bildung von Verbindungen des Bors und Siliciums mit Stickstoff und gewissen Metallen. J. Prakt. Chem. 1842, 27, 422–430.
  15. Chopra, N.G.; Luyken, R.J.; Cherrey, K.; Crespi, V.H.; Cohen, M.L.; Louie, S.G.; Zettl, A. Boron Nitride Nanotubes. Science 1995, 269, 966–967.
  16. Yu, I.; Jo, Y.; Ko, J.; Moon, S.Y.; Ahn, S.Y.M.; Joo, Y. Highly Aligned Array of Heterostructured Polyflourene-Isolated Boron Nitride and Carbon Nanotubes. ACS Appl. Mater. Interfaces 2021, 13, 12417–12424.
  17. Kovalskii, A.M.; Matveev, A.T.; Lebedev, O.I.; Sukhorukova, I.V.; Firestein, K.L.; Steinman, A.E.; Shtansky, D.V.; Golberg, D. Growth of Spherical Boron Oxynitride Banoparticles with Smooth and Petalled Surfaces during Chemical Vapor Deposition Process. CrystEngComm 2016, 18, 6689–6699.
  18. Le, T.-H.; Oh, Y.; Kim, H.; Yoon, H. Exfoliation of 2D Materials for Energy and Environmental Applications. Chem. Eur. J. 2022, 26, 6360–6401.
  19. Gautam, C.; Chelliah, S. Methods of Hexagonal Boron Nitride Exfoliation and Its Functionalization: Covalent and Non-Covalent Approaches. RSC Adv. 2021, 11, 31284–31327.
  20. Juma, I.G.; Kim, G.; Jariwala, D.; Behura, S.K. Direct Growth of Hexagonal Boron Nitride on Non-Metallic Substrates and Its Heterostructures with Graphene. iScience 2021, 24, 103374–103393.
  21. Shen, X.; Zheng, Q.; Kim, J.-K. Rational Design of Two-Dimensional Nanofillers for Polymer Nanocomposites toward Multifunctional Applications. Prog. Mater. Sci. 2021, 115, 100708–100773.
  22. Meziani, M.J.; Sheriff, K.; Parajuli, P.; Priego, P.; Bhattacharya, S.; Rao, A.M.; Quimby, J.L.; Qiao, R.; Wang, P.; Hwu, S.-J.; et al. Advances in Studies of Boron Nitride Nanosheets and Nanocomposites for Thermal Transport and Related Applications. ChemPhysChem 2022, 23, e202100645–e202100668.
  23. Revabhai, P.M.; Singhal, R.K.; Basu, H.; Kailasa, S.K. Progress on Boron Nitride Nanostructure Materials: Properties, Synthesis and Applications in Hydrogen Storage and Analytical Chemistry. J. Nanostruct. Chem. 2022.
  24. Roy, S.; Zhang, X.; Puthirath, A.B.; Meiyazhagan, A.; Bhattacharyya, S.; Rahman, M.M.; Babu, G.; Susarla, S.; Saju, S.K.; Tran, M.K.; et al. Structure, Properties and Applications of Two-Dimensional Hexagonal Boron Nitride. Adv. Mater. 2021, 33, 2101589.
  25. Yang, Y.; Peng, Y.; Saleem, M.F.; Chen, Z.; Sun, W. Hexagonal Boron Nitride on III–V Compounds: A Review of the Synthesis and Applications. Materials 2022, 15, 4396.
  26. Yasnó, J.; Kiminami, R.H.G.A. Short Time Reaction Synthesis of Nano-Hexagonal Boron Nitride. Adv. Powder Technol. 2020, 31, 4436–4443.
  27. Matveev, A.T.; Permyakova, E.S.; Kovalskii, A.M.; Leibo, D.; Shchetinin, I.V.; Maslakov, K.I.; Golberg, D.V.; Shtansky, D.V.; Konopatsky, A.S. New Insights into Synthesis of Nanocrystalline Hexagonal BN. Ceram. Int. 2020, 46, 19866–19872.
  28. Köken, D.; Sungur, P.; Cebeci, H.; Cebeci, F.C. Revealing the Effect of Sulfur Compounds for Low-Temperature Synthesis of Boron Nitride Nanotubes from Boron Minerals. ACS Appl. Nano Mater. 2022, 5, 2137–2146.
  29. Silva-Santos, S.D.; Impellizzeri, A.; Aguiar, A.L.; Journet, C.; Dalverny, C.; Toury, B.; De Sousa, J.M.; Ewels, C.P.; San-Miguel, A. High Pressure in Boron Nitride Nanotubes for Kirigami Nanoribbon Elaboration. J. Phys. Chem. C 2021, 125, 11440–11453.
  30. Pham, T.; Stonemeyer, S.; Marquez, J.; Long, H.; Gillbert, M.; Worsley, M.; Zettl, A. One-Step Conversion of Graphite to Crinkled Boron Nitride Nanofoams for Hydrophobic Liquid Absorption. ACS Appl. Nano Mater. 2021, 4, 3500–3507.
  31. Matsoso, B.; Vuillet-a-Ciles, V.; Bois, L.; Toury, B.; Journet, C. Improving Formation Conditions and Properties of hBN Nanosheets through BaF2-Assisted Polymer Derived Ceramics (PDCs) Technique. Nanomaterials 2020, 10, 443.
  32. Liu, F.; Han, R.; Naficy, S.; Casillas, G.; Sun, X.; Huang, Z. Few-Layered Boron Nitride Nanosheets for Strengthening Polyurethane Hydrogels. ACS Appl. Nano Mater. 2021, 4, 7988–7994.
  33. Kondo, D.; Kataoka, M.; Hayashi, K.; Sato, S. Few-Layer Hexagonal Boron Nitride Synthesized by Chemical Vapor Deposition and Its Insulating Properties. Nano Express 2021, 2, 030001–030007.
  34. Shi, Z.; Wang, X.; Li, Q.; Yang, P.; Lu, G.; Jiang, R.; Wang, H.; Zhang, C.; Cong, C.; Liu, Z.; et al. Vapor–Liquid–Solid Growth of Large-Area Multilayer Hexagonal Boron Nitride on Dielectric Substrates. Nat. Comm. 2020, 11, 849–857.
  35. Li, Y.; Garnier, V.; Steyer, P.; Journet, C.; Toury, B. Millimeter-Scale Hexagonal Boron Nitride Single Crystals for Nanosheet Generation. ACS Appl. Nano Mater. 2020, 3, 1508–1515.
  36. Li, J.; Wang, J.; Zhang, X.; Elias, C.; Ye, G.; Evans, D.; Eda, G.; Redwing, J.M.; Cassabois, G.; Gil, B.; et al. Hexagonal Boron Nitride Crystal Growth from Iron, a Single Component Flux. ACS Nano 2021, 15, 7032–7039.
  37. Bansal, A.; Hilse, M.; Huet, B.; Wang, K.; Kozhakhmetov, A.; Kim, J.H.; Bachu, S.; Alem, N.; Collazo, R.; Robinson, J.A.; et al. Substrate Modification During Chemical Vapor Deposition of hBN on Sapphire. ACS Appl. Mater. Interfaces 2021, 13, 54516–54526.
  38. Chen, J.; Wang, G.; Meng, J.; Cheng, Y.; Yin, Z.; Tian, Y.; Huang, J.; Zhang, S.; Wu, J.; Zhang, X. Low-Temperature Direct Growth of Few-Layer Hexagonal Boron Nitride on Catalyst-Free Sapphire Substrates. ACS Appl. Mater. Interfaces 2022, 14, 7004–7011.
  39. Singhal, R.; Echeverria, E.; McIlroy, D.N.; Singh, R.N. Synthesis of Hexagonal Boron Nitride Films on Silicon and Sapphire Substrates by Low-Pressure Chemical Vapor Deposition. Thin Solid Film. 2021, 733, 138812.
  40. Gigliotti, J.; Li, X.; Sundaram, S.; Deniz, D.; Prudkovskiy, V.; Turmaud, J.-P.; Hu, Y.; Hu, Y.; Fossard, F.; Meérot, J.-S.; et al. Highly Ordered Boron Nitride/Epigraphene Epitaxial Films on Silicon Carbide by Lateral Epitaxial Deposition. ACS Nano 2020, 14, 12962–12971.
  41. Arjmandi-Tash, H. In Situ Growth of Graphene on Hexagonal Boron Nitride for Electronic Transport Applications. J. Mater. Chem. C 2020, 8, 380–386.
  42. Yan, Q.; Dai, W.; Gao, J.; Tan, X.; Lv, L.; Ying, J.; Lu, X.; Lu, J.; Yao, Y.; Wei, Q.; et al. Ultrahigh-Aspect-Ratio Boron Nitride Nanosheets Leading to Superhigh In-Plane Thermal Conductivity of Foldable Heat Spreader. ACS Nano 2021, 15, 6489–6498.
  43. Li, S.; Lu, X.; Lou, Y.; Liu, K.; Zou, B. The Synthesis and Characterization of h-BN Nanosheets with High Yield and Crystallinity. ACS Omega 2021, 6, 27814–27822.
  44. Wang, Z.-G.; Wei, X.; Bai, M.-H.; Lei, J.; Xu, L.; Huang, H.-D.; Du, J.; Dai, K.; Xu, J.-Z.; Li, Z.-M. Green Production of Covalently Functionalized Boron Nitride Nanosheets Via Saccharide-Assisted Mechanochemical Exfoliation. ACS Sustain. Chem. Eng. 2021, 9, 11155–11162.
  45. Huang, J.; E, S.; Li, J.; Jia, F.; Ma, Q.; Hua, L.; Lu, Z. Ball-Milling Exfoliation of Hexagonal Boron Nitride in Viscous Hydroxyethyl Cellulose for Producing Nanosheet Films as Thermal Interface Materials. ACS Appl. Nano Mater. 2021, 4, 13167–13175.
  46. Yusupov, K.U.; Corthay, S.; Bondarev, A.V.; Kovalskii, A.M.; Matveev, A.T.; Arkhipov, D.; Golberg, D.; Shtansky, D.V. Spark Plasma Sintered Al-based Composites Reinforced with BN Nanosheets Exfoliated under Ball Milling in Ethylene Glycol. Mater. Sci. Eng. A 2019, 745, 74–81.
  47. Ding, J.-H.; Zhao, H.-R.; Yu, H.-B. High-Yield Synthesis of Extremely High Concentrated and Few-Layered Boron Nitride Nanosheet Dispersions. 2D Mater. 2018, 5, 045015–045033.
  48. Du, Y.; Zhang, Y.; Zhang, R.; Lin, S. Synthesis of Ultrathin Functional Boron Nitride Nanosheets and Their Application in Anticorrosion. ACS Appl. Nano Mater. 2021, 4, 11088–11096.
  49. Kovalskii, A.M.; Matveev, A.T.; Popov, Z.I.; Volkov, I.N.; Sukhanova, E.V.; Lytkina, A.A.; Yaroslavtsev, A.B.; Konopatsky, A.S.; Leybo, D.V.; Bondarev, A.V.; et al. (Ni,Cu)/Hexagonal BN Nanohybrids—New Efficient Catalysts for Methanol Steam Reforming and Carbon Monoxide Oxidation. Chem. Eng. J. 2020, 395, 125109.
  50. Konopatsky, A.S.; Firestein, K.L.; Leybo, D.V.; Sukhanova, E.V.; Popov, Z.I.; Fang, X.; Manakhov, A.M.; Kovalskii, A.M.; Matveev, A.T.; Shtansky, D.V.; et al. Structural Evolution of Ag/BN Hybrids via a Polyol-Assisted Fabrication Process and Their Catalytic Activity in CO oxidation. Catal. Sci. Technol. 2019, 9, 6460–6470.
  51. Grant, J.T.; Carrero, C.A.; Goeltl, F.; Venegas, J.; Mueller, P.; Burt, S.P.; Specht, S.E.; McDermott, W.P.; Chieregato, A.; Hermans, I. Selective Oxidative Dehydrogenation of Propane to Propene Using Boron Nitride Catalysts. Science 2016, 354, 1570–1573.
  52. Venegas, J.M.; Hermans, I. The Influence of Reactor Parameters on the Boron Nitride-Catalyzed Oxidative Dehydrogenation of Propane. Org. Process Res. Dev. 2018, 22, 1644–1652.
  53. McDermott, W.P.; Venegas, J.; Hermans, I. Selective Oxidative Cracking of N-butane to Light Olefins over Hexagonal Boron Nitride with Limited Formation of COx. ChemSusChem 2020, 13, 152–158.
  54. Kraus, P.; Lindstedt, R.P. It’s a Gas: Oxidative Dehydrogenation of Propane over Boron Nitride Catalysts. J. Phys. Chem. C 2021, 125, 5623–5634.
  55. Du, M.; Liu, Q.; Huang, C.; Qiu, X. One-Step Synthesis of Magnetically Recyclable Core–Shell Nanocatalysts for Catalytic Reduction of Nitroarenes. RSC Adv. 2017, 7, 35451–35459.
  56. Chen, N.; Zhu, Z.; Su, T.; Liao, W.; Deng, C.; Ren, W.; Zhao, Y.; Lü, H. Catalytic Hydrogenolysis of Hydroxymethylfurfural to Highly Selective 2,5-Dimethylfuran over FeCoNi/h-BN Catalyst. Chem. Eng. J. 2020, 381, 122755.
  57. Zhang, R.; Wang, L.; Yang, X.; Tao, Z.; Ren, X.; Lv, B. The Role of Surface NH Groups on the Selective Hydrogenation of Cinnamaldehyde over Co/BN Catalysts. Appl. Surf. Sci. 2019, 492, 736–745.
  58. Ji, H.; Ju, H.; Lan, R.; Wu, P.; Sun, J.; Chao, Y.; Xun, S.; Zhu, W.; Li, H. Phosphomolybdic Acid Immobilized on Ionic Liquid-Modified Hexagonal Boron Nitride for Oxidative Desulfurization of Fuel. RSC Adv. 2017, 7, 54266–54276.
  59. Cheng, S.-Q.; Weng, X.-F.; Wang, Q.-N.; Zhou, B.-C.; Li, W.-C.; Li, M.-R.; He, L.; Wang, D.-Q.; Lu, A.-H. Defect-Rich BN-Supported Cu with Superior Dispersion for Ethanol Conversion to Aldehyde and Hydrogen. Chin. J. Catal. 2022, 43, 1092–1100.
  60. Konopatsky, A.S.; Leybo, D.V.; Firestein, K.L.; Chepkasov, I.V.; Popov, Z.I.; Permyakova, E.S.; Volkov, I.N.; Kovalskii, A.M.; Matveev, A.T.; Shtansky, D.V.; et al. Polyol Synthesis of Ag/BN Nanohybrids and Their Catalytic Stability in CO Oxidation Reaction. ChemCatChem 2020, 12, 1691–1698.
  61. Wang, L.; Wang, Y.; Zhang, C.-W.; Wen, J.; Weng, X.; Shi, L. A Boron Nitride Nanosheet-Supported Pt/Cu Cluster as a High-Efficiency Catalyst for Propane Dehydrogenation. Catal. Sci. Technol. 2020, 10, 1248–1255.
  62. Li, L.; Liu, X.; He, H.; Zhang, N.; Liu, Z.; Zhang, G. A Novel Two-Dimensional MgO-h-BN Nanomaterial Supported Pd Catalyst for CO Oxidation Reaction. Catal. Today 2019, 332, 214–221.
  63. Lu, M.; Zhang, X.; Deng, J.; Kuboon, S.; Faungnawakij, K.; Xiao, S.; Zhang, D. Coking-Resistant Dry Reforming of Methane over BN–Nanoceria Interface-Confined Ni Catalysts. Catal. Sci. Technol. 2020, 10, 4237–4244.
  64. Kovalskii, A.M.; Volkov, I.N.; Evdokimenko, N.D.; Leybo, D.V.; Chepkasov, I.V.; Popov, Z.I.; Matveev, A.T.; Manahov, A.M.; Permyakova, E.S.; Konopatsky, A.S.; et al. (Au and Pt)/Hexagonal BN Nanohybrids in Carbon Monoxide Oxidation and Carbon Dioxide Hydrogenation Reactions. Appl. Catal. B-Environ. 2022, 303, 120891.
  65. Konopatsky, A.S.; Firestein, K.L.; Evdokimenko, N.D.; Baidyshev, V.S.; Chepkasov, I.V.; Popov, Z.I.; Matveev, A.T.; Shetinin, I.V.; Leybo, D.V.; Volkov, I.N.; et al. Microstructure and Catalytic Properties of Fe3O4/BN, Fe3O4(Pt)/BN, and FePt/BN Nanohybrids in CO2 Hydrogenation Reaction: Experimental and Theoretical Insights. J. Catal. 2021, 402, 130–142.
  66. Sheng, J.; Li, W.-C.; Lu, W.-D.; Yan, B.; Qiu, B.; Gao, X.-Q.; Zhang, R.-P.; Zhou, S.-Z.; Lu, A.-H. Preparation of Oxygen Reactivity-Tuned FeOx/BN Catalyst for Selectively Oxidative Dehydrogenation of Ethylbenzene to Styrene. Appl. Catal. B-Environ. 2022, 305, 121070.
  67. Wang, Y.; Wang, J.; Zheng, P.; Sun, C.; Luo, J.; Xie, X. Boosting Selectivity and Stability on Pt/BN Catalysts for Propane Dehydrogenation via Calcination & Reduction-Mediated Strong Metal-Support Interaction. J. Energy Chem. 2022, 67, 451–457.
  68. Wang, X.; Zhang, C.; Chang, Q.; Wang, L.; Lv, B.; Xu, J.; Xiang, H.; Yang, Y.; Li, Y. Enhanced Fischer-Tropsch Synthesis Performances of Fe/h-BN Catalysts by Cu and Mn. Catal. Today 2020, 343, 91–100.
  69. Peinado, C.; Liuzzi, D.; Ladera-Gallardo, R.M.; Retuerto, M.; Ojeda, M.; Peña, M.A.; Rojas, S. Effects of Support and Reaction Pressure for the Synthesis of Dimethyl Ether over Heteropolyacid Catalysts. Sci. Rep. 2020, 10, 8551.
  70. Li, X.; Lin, B.; Li, H.; Yu, Q.; Ge, Y.; Jin, X.; Liu, X.; Zhou, Y.; Xiao, J. Carbon Doped Hexagonal BN as a Highly Efficient Metal-Free Base Catalyst for Knoevenagel Condensation Reaction. Appl. Catal. B-Environ. 2018, 239, 254–259.
  71. Wang, X.; Liu, Z.; Shi, X.; Jia, Y.; Zhu, G.; Peng, J.; Wang, Q. Optical and Photocatalytic Characteristics of Se-Doped h-Boron Nitride: Experimental Assessments and DFT Calculations. J. Alloys Compd. 2022, 909, 164791.
  72. Wang, J.; Sun, C.; Xia, W.; Cao, Z.; Sheng, G.; Xie, X. Pd/BN Catalysts for Highly Efficient Hydrogenation of Maleic Anhydride to Succinic Anhydride. Appl. Catal. A-Gen. 2022, 630, 118471.
  73. Wang, Y.; Chen, J.; Wang, L.; Weng, H.; Wu, Z.; Jiao, L.; Muroya, Y.; Yamashita, S.; Cheng, S.; Li, F.; et al. γ-Radiation Synthesis of Ultrasmall Noble Metal (Pd, Au, Pt) Nanoparticles Embedded on Boron Nitride Nanosheets for High-Performance Catalysis. Ceram. Int. 2021, 47, 26963–26970.
  74. Cao, Z.; Bu, J.; Zhong, Z.; Sun, C.; Zhang, Q.; Wang, J.; Chen, S.; Xie, X. Selective Hydrogenation of Cinnamaldehyde to Cinnamyl Alcohol over BN-Supported Pt Catalysts at Room Temperature. Appl. Catal. A-Gen. 2019, 578, 105–115.
  75. Gao, M.; Jia, X.; Ma, J.; Fan, X.; Gao, J.; Xu, J. Self-Regulated Catalysis for the Selective Synthesis of Primary Amines from Carbonyl Compounds. Green Chem. 2021, 23, 7115–7121.
  76. Zhu, W.; Wu, Z.; Foo, G.S.; Gao, X.; Zhou, M.; Liu, B.; Veith, G.M.; Wu, P.; Browning, K.L.; Lee, H.N.; et al. Taming Interfacial Electronic Properties of Platinum Nanoparticles on Vacancy-Abundant Boron Nitride Nanosheets for Enhanced Catalysis. Nat. Comm. 2017, 8, 15291.
  77. Lu, L.-J.; Wu, P.-W.; He, J.; Hua, M.-Q.; Chao, Y.-H.; Yang, N.; Chen, L.-L.; Jiang, W.; Fan, L.; Ji, H.-B.; et al. N-Hydroxyphthalimide Anchored on Hexagonal Boron Nitride as a Metal-Free Heterogeneous Catalyst for Deep Oxidative Desulfurization. Pet. Sci. 2021, 19, 1382–1389.
  78. Rana, P.; Dixit, R.; Sharma, S.; Dutta, S.; Yadav, S.; Sharma, A.; Kaushik, B.; Rana, P.; Adholeya, A.; Sharma, R.K. Enhanced Catalysis through Structurally Modified Hybrid 2-D Boron Nitride Nanosheets Comprising of Complexed 2-Hydroxy-4-Methoxybenzophenone Motif. Sci. Rep. 2021, 11, 24429.
  79. Wang, A.; Li, J.; Zhang, T. Heterogeneous Single-Atom Catalysts. Nat. Rev. Chem. 2018, 2, 65–81.
  80. Kaiser, S.K.; Chen, Z.; Akl, D.F.; Mitchell, S.; Pérez-Ramírez, J. Single-Atom Catalysts across the Periodic Table. Chem. Rev. 2020, 120, 11703–11809.
  81. Lei, Y.; Pakhira, S.; Fujisawa, K.; Liu, H.; Guerrero-Bermea, C.; Zhang, T.; Dasgupta, A.; Martinez, L.M.; Singamaneni, S.R.; Wang, K.; et al. Low Temperature Activation of Inert Hexagonal Boron Nitride for Metal Deposition and Single Atom Catalysis. Mater. Today 2021, 51, 108–116.
  82. Li, Z.; Wei, W.; Li, H.; Li, S.; Leng, L.; Zhang, M.; Horton, J.H.; Wang, D.; Sun, W.; Guo, C.; et al. Low-Temperature Synthesis of Single Palladium Atoms Supported on Defective Hexagonal Boron Nitride Nanosheet for Chemoselective Hydrogenation of Cinnamaldehyde. ACS Nano 2021, 15, 10175–10184.
  83. Huang, X.; Akdim, O.; Douthwaite, M.; Wang, K.; Zhao, L.; Lewis, R.J.; Pattisson, S.; Daniel, I.T.; Miedziak, P.J.; Shaw, G.; et al. Au–Pd Separation Enhances Bimetallic Catalysis of Alcohol Oxidation. Nature 2022, 603, 271–275.
  84. Zhu, J.; Xu, W.; Chen, J.; Gan, Z.; Wang, X.; Zhou, J. Development of Core–Shell Structured Mo2 as Novel Microwave Catalysts for Highly Effective Direct Decomposition of H2S into H2 and S at Low Temperature. Catal. Sci. Technol. 2020, 10, 6769–6779.
  85. Erusappan, E.; Thiripuranthagan, S.; Durai, M.; Kumaravel, S.; Vembuli, T. Photocatalytic Performance of Visible Active Boron Nitride Supported ZnFe2O4 (ZnFe2O4/BN) Nanocomposites for the Removal of Aqueous Organic Pollutants. New J. Chem. 2020, 44, 7758–7770.
  86. Ren, K.; Dong, Y.; Chen, Y.; Shi, H. Bi2WO6 Nanosheets Assembled BN Quantum Dots: Enhanced Charge Separation and Photocatalytic Antibiotics Degradation. Colloids Surf. A 2022, 637, 128208.
  87. Yan, T.; Du, Z.; Wang, J.; Cai, H.; Bi, D.; Guo, Z.; Liu, Z.; Tang, C.; Fang, Y. Construction of 2D/2D Bi2WO6/BN Heterojunction for Effective Improvement on Photocatalytic Degradation of Tetracycline. J. Alloys Compd. 2022, 894, 162487.
  88. Li, W.; Wang, F.; Chu, X.; Dang, Y.; Liu, X.; Ma, T.; Li, J.; Wang, C. 3D Porous BN/RGO Skeleton Embedded by MoS2 Nanostructures for Simulated-Solar-Light Induced Hydrogen Production. Chem. Eng. J. 2022, 435, 132441.
  89. Jiang, C.; Zhang, M.; Dong, G.; Wei, T.; Feng, J.; Ren, Y.; Luan, T. Photocatalytic Nitrate Reduction by a Non-Metal Catalyst h-BN: Performance and Mechanism. Chem. Eng. J. 2022, 429, 132216.
  90. Haroon, H.; Wahid, M.; Majid, K. Metal−Organic Framework-Derived p-Type Cu3P/Hexagonal Boron Nitride Nanostructures for Photocatalytic Oxidative Coupling of Aryl Halides to Biphenyl Derivatives. ACS Appl. Nano Mater. 2022, 5, 2006–2017.
  91. Bao, Y.; Yan, W.; Sun, P.-P.; Seow, J.Z.Y.; Lua, S.K.; Lee, W.J.; Liang, Y.N.; Lim, T.-T.; Xu, Z.J.; Zhou, K.; et al. Unexpected Intrinsic Catalytic Function of Porous Boron Nitride Nanorods for Highly Efficient Peroxymonosulfate Activation in Water Treatment. ACS Appl. Mater. Interfaces 2022, 14, 18409–18419.
  92. Matveev, A.T.; Konopatsky, A.S.; Leybo, D.V.; Volkov, I.N.; Kovalskii, A.M.; Varlamova, L.A.; Sorokin, P.B.; Fang, X.; Shtansky, D.V. Amorphous MoSxOy/h-BNxOy Nanohybrids: Synthesis and Dye Photodegradation. Nanomaterials 2021, 11, 3232.
  93. Zhao, G.; Wang, A.; He, W.; Xing, Y.; Xu, X. 2D New Nonmetal Photocatalyst of Sulfur-Doped h-BN Nanosheeets with High Photocatalytic Activity. Adv. Mater. Interfaces 2019, 6, 1900062.
  94. Sun, M.; Lv, Y.; Song, Y.; Wu, H.; Wang, G.; Zhang, H.; Chen, M.; Fu, Q.; Bao, X. CO-Tolerant /C Core–Shell Electrocatalysts for Proton Exchange Membrane Fuel Cells. Appl. Surf. Sci. 2018, 450, 244–250.
  95. Yesudoss, D.K.; Lee, G.; Shanmugam, S. Strong Catalyst Support Interactions in Defect-Rich γ-Mo2N Nanoparticles Loaded 2D-h-BN Hybrid for Highly Selective Nitrogen Reduction Reaction. Appl. Catal. B-Environ. 2021, 287, 119952.
  96. Min, Y.; Zhou, X.; Chen, J.-J.; Chen, W.; Zhou, F.; Wang, Z.; Yang, J.; Xiong, C.; Wang, Y.; Li, F.; et al. Integrating Single-Cobalt-Site and Electric Field of Boron Nitride in Dechlorination Electrocatalysts by Bioinspired Design. Nat. Comm. 2021, 12, 303.
  97. Balta, Z.; Simsek, E.B. Understanding the Structural and Photocatalytic Effects of Incorporation of Hexagonal Boron Nitride Whiskers into Ferrite Type Perovskites (BiFeO3, MnFeO3) for Effective Removal of Pharmaceuticals from Real Wastewater. J. Alloys Compd. 2022, 898, 162897.
  98. Merlo, A.; Mokkapati, V.R.S.S.; Pandit, S.; Mijakovic, I. Boron Nitride Nanomaterials: Biocompatibility and Bio-applications. Biomater. Sci. 2018, 6, 2298–2311.
  99. Sen, O.; Melis, E.; Çulha, M. One-Step Synthesis of Hexagonal Boron Nitrides, Their Crystallinity and Biodegradation. Front. Bioeng. Biotechnol. 2018, 6, 83.
  100. Şen, Ö.; Emanet, M.; Çulha, M. Stimulatory Effect of Hexagonal Boron Nitrides in Wound Healing. ACS Appl. Bio Mater. 2019, 2, 5582–5596.
  101. Sogut, I.; Paltun, S.O.; Tuncdemir, M.; Ersoz, M.; Hurdag, C. The Antioxidant and Antiapoptotic Effect of Boric Acid on Hepatoxicity in Chronic Alcohol-Fed Rats. Can. J. Physiol. Pharmacol. 2018, 96, 404–411.
  102. Donbaloglu, F.; Leblebici, S. Interactive Effect of Boric Acid and Temperature Stress on Phenological Characteristics and Antioxidant System in Helianthus annuus L. South Afr. J. Bot. 2022, 147, 391–399.
  103. Ince, S.; Kucukkurt, I.; Cigerci, H.I.; Fidan, A.F.; Eryavuz, A. The Effects of Dietary Boric Acid and Borax Supplementation on Lipid Peroxidation, Antioxidant Activity, and DNA Damage in Rats. J. Trace Elem. Med. Biol. 2010, 24, 161–164.
  104. Song, S.; Gao, P.; Sun, L.; Kang, D.; Kongsted, J.; Poongavanam, V.; Zhan, P.; Liu, X. Recent Developments in the Medicinal Chemistry of Single Boron Atom-Containing Compounds. Acta Pharm. Sin. B 2021, 11, 3035–3059.
  105. Taskin, I.C.; Sen, O.; Emanet, M.; Culha, M. Hexagonal Boron Nitrides Reduce the Oxidative Stress on Cells. Nanotechnology 2020, 31, 215101.
  106. Kar, F.; Söğüt, I.; Hacioğlu, C. Hexagonal Boron Nitride Nanoparticles Trigger Oxidative Stress by Modulating Thiol/Disulfide Homeostasis. Hum. Exp. Toxicol. 2021, 40, 1572–1583.
  107. Yegin, B.; Ozkazanc, H.; Er, K.D.; Ozkazanc, E. Antimicrobial Performance and Charge Transport Mechanism of Poly(N-Methylpyrrole)-Boron Nitride Composite. Mater. Chem. Phys. 2022, 278, 125709.
  108. Pandit, S.; Gaska, K.; Mokkapati, V.R.S.S.; Forsberg, S.; Svensson, M.; Kádár, R.; Mijakovic, I. Antibacterial effect of boron nitride flakes with controlled orientation in polymer composites. RSC Adv. 2019, 9, 33454–33459.
  109. Mukheem, A.; Shahabuddin, S.; Akbar, N.; Miskon, A.; Sarih, N.M.; Sudesh, K.; Khan, N.A.; Saidur, R.; Sridewi, N. Boron Nitride Doped Polyhydroxyalkanoate/Chitosan Nanocomposite for Antibacterial and Biological Applications. Nanomaterials 2019, 9, 645.
  110. Xiong, S.; Fu, P.; Zou, Q.; Chen, L.; Jiang, M.; Zhang, P.; Wang, Z.; Cui, L.; Guo, H.; Gai, J. Heat Conduction and Antibacterial Hexagonal Boron Nitride/Polypropylene Nanocomposite Fibrous Membranes for Face Masks with Long-Time Wearing Performance. ACS Appl. Mater. Interfaces 2020, 13, 196–206.
  111. Onyszko, M.; Markowska-Szczupak, A.; Rakoczy, R.; Paszkiewicz, O.; Janusz, J.; Gordon-Kuza, A.; Wenelska, K.; Mijowska, E. Few Layered Oxidized h-BN as Nanofiller of Cellulose-Based Paper with Superior Antibacterial Response and Enhanced Mechanical/Thermal Performance. Int. J. Mol. Sci. 2020, 21, 5396.
  112. Gudz, K.Y.; Antipina, L.Y.; Permyakova, E.S.; Kovalskii, A.M.; Konopatsky, A.S.; Filippovich, S.Y.; Dyatlov, I.A.; Slukin, P.V.; Ignatov, S.G.; Shtansky, D.V. Ag-Doped and Antibiotic-Loaded Hexagonal Boron Nitride Nanoparticles as Promising Carriers to Fight Different Pathogens. ACS Appl. Mater. Interfaces 2021, 13, 23452–23468.
  113. Kivanç, M.; Barutca, B.; Koparal, A.T.; Göncü, Y.; Bostancı, S.H.; Ay, N. Effects of Hexagonal Boron Nitride Nanoparticles on Antimicrobial and Antibiofilm Activities, Cell Viability. Mater. Sci. Eng. C 2018, 91, 115–124.
  114. Sert, B.; Gonca, S.; Ozay, Y.; Harputlu, E.; Ozdemir, S.; Ocakoglu, K. Biointerfaces Investigation of the Antifouling Properties of Polyethersulfone Ultrafiltration Membranes by Blending of Boron Nitride Quantum Dots. Colloid Surf. B 2021, 205, 111867.
  115. Zhang, Y.; Chan, C.; Li, Z.; Ma, J.; Meng, Q.; Zhi, C.; Sun, H.; Fan, J. Nanotoxicity of Boron Nitride Nanosheet to Bacterial Membranes. Langmuir 2019, 35, 6179–6187.
  116. Gudz, K.Y.; Permyakova, E.S.; Matveev, A.T.; Bondarev, A.V.; Manakhov, A.M.; Sidorenko, D.A.; Filippovich, S.Y.; Brouchkov, A.V.; Golberg, D.V.; Ignatov, S.G.; et al. Pristine and Antibiotic-Loaded Nanosheets/Nanoneedles-Based Boron Nitride Films as a Promising Platform to Suppress Bacterial and Fungal Infections. ACS Appl. Mater. Interfaces 2020, 12, 42485–42498.
  117. Rasel, M.A.I.; Li, T.; Nguyen, T.D.; Singh, S.; Zhou, Y.; Gu, Y.T. Biophysical Response of Living Cells to Boron Nitride Nanoparticles: Uptake Mechanism and Bio-Mechanical Characterization. J. Nanoparticle Res. 2015, 17, 441.
  118. Li, X.; Wang, X.; Zhang, J.; Hanagata, N.; Wang, X.; Weng, Q.; Ito, A.; Bando, Y.; Golberg, D. Hollow Boron Nitride Nanospheres as Boron Reservoir for Prostate Cancer Treatment. Nat. Commun. 2017, 8, 13936.
  119. Volkmann, M.; Meyns, M.; Lesyuk, R.; Lehmann, H.; Klinke, C. Attachment of Colloidal Nanoparticles to Boron Nitride Nanotubes. Chem. Mater. 2017, 29, 726–734.
  120. Ikram, M.; Hussain, I.; Hassan, J.; Haider, A.; Imran, M.; Aqeel, M.; Ul-hamid, A.; Ali, S. Evaluation of Antibacterial and Catalytic Potential of Copper-Doped Chemically Exfoliated Boron Nitride Nanosheets. Ceram. Int. 2020, 46, 21073–21083.
  121. Ikram, M.; Jahan, I.; Haider, A.; Hassan, J.; Ul-Hamid, A.; Imran, M.; Haider, J.; Shahzadi, A.; Shahbaz, A.; Ali, S. Bactericidal Behavior of Chemically Exfoliated Boron Nitride Nanosheets Doped with Zirconium. Appl. Nanosci. 2020, 10, 2339–2349.
  122. Pan, Y.; Zheng, H.; Li, G.; Li, Y.; Jiang, J.; Chen, J.; Xie, Q.; Wu, D.; Ma, R.; Liu, X.; et al. Antibiotic-Like Activity of Atomic Layer Boron Nitride for Combating Resistant Bacteria. ACS Nano 2022, 16, 7674–7688.
  123. Xie, X.; Hou, Z.; Duan, G.; Zhang, S.; Zhou, H.; Yang, Z.; Zhou, R. Boron Nitride Nanosheets Elicit Significant Hemolytic Activity via Destruction of Red Blood Cell Membranes. Colloid Surf. B 2021, 203, 111765.
  124. Mapleback, B.J.; Brack, N.; Thomson, L.; Spencer, M.J.S.; Osborne, D.A.; Doshi, S.; Thostenson, E.T.; Rider, A.N. Development of Stable Boron Nitride Nanotube and Hexagonal Boron Nitride Dispersions for Electrophoretic Deposition. Langmuir 2020, 36, 3425–3438.
  125. Czarniewska, E.; Mrówczyńska, L.; Jędrzejczak-Silicka, M.; Nowicki, P.; Trukawka, M.; Mijowska, E. Non-Cytotoxic Hydroxyl-Functionalized Exfoliated Boron Nitride Nanoflakes Impair the Immunological Function of Insect Haemocytes in Vivo. Sci. Rep. 2019, 9, 14027.
  126. Maharjan, S.; Gautam, M.; Poudel, K.; Yong, C.S.; Ku, S.K.; Kim, J.O.; Byeon, J.H. Biomaterials Streamlined Plug-in Aerosol Prototype for Reconfigurable Manufacture of Nano-Drug Delivery Systems. Biomaterials 2022, 284, 121511.
  127. Pasquale, D.D.; Marino, A.; Tapeinos, C.; Pucci, C. Homotypic Targeting and Drug Delivery in Glioblastoma Cells Through Cell Membrane-Coated Boron Nitride Nanotubes. Mater. Des. 2020, 192, 108742.
  128. Liu, J.; Zheng, T.; Tian, Y. Photodynamic Therapy Functionalized h-BN Nanosheets as a Theranostic Platform for SERS Real-Time Monitoring of MicroRNA and Photodynamic Therapy Angewandte. Angew. Chem. Int. Ed. 2019, 58, 7757–7761.
  129. Ahmad, P.; Khandaker, M.U.; Muhammad, N.; Rehman, F.; Ullah, Z.; Khan, G.; Khan, M.I.; Haq, S.; Ali, H.; Khan, A.; et al. Synthesis of Enriched Boron Nitride Nanocrystals: A Potential Element for Biomedical Applications. Appl. Radiat. Isot. 2020, 166, 109404.
  130. Malouff, T.D.; Seneviratne, D.S.; Ebner, D.K.; Stross, W.C.; Waddle, M.R.; Trifiletti, D.M.; Krishnan, S. Boron Neutron Capture Therapy: A Review of Clinical Applications. Front. Oncol. 2021, 11, 601820.
  131. Yokoyama, K.; Miyatake, S.-I.; Kajimoto, Y.; Kawabata, S.; Doi, A.; Yoshida, T.; Asano, T.; Kirihata, M.; Ono, K.; Kuroiwa, T. Pharmacokinetic Study of BSH and BPA in Simultaneous Use for BNCT. J. Neurooncol. 2006, 78, 227–232.
  132. Li, L.; Li, J.; Shi, Y.; Du, P.; Zhang, Z.; Liu, T.; Zhang, R.; Liu, Z. On-Demand Biodegradable Boron Nitride Nanoparticles for Treating Triple Negative Breast Cancer with Boron Neutron Capture Therapy. ACS Nano 2019, 13, 13843–13852.
  133. Li, L.; Dai, K.; Li, J.; Shi, Y.; Zhang, Z.; Liu, T.; Xie, J.; Zhang, R.; Liu, Z. A Boron-10 Nitride Nanosheet for Combinational Boron Neutron Capture Therapy and Chemotherapy of Tumor. Biomaterials 2021, 268, 120587.
  134. Silva, W.M.; Ribeiro, H.; Taha-Tijerina, J.J. Potential Production of Theranostic Boron Nitride Nanotubes (64 Cu-BNNTs) Radiolabeled by Neutron Capture. Nanomaterials 2021, 11, 2907.
  135. Shuai, C.; Gao, C.; Feng, P.; Xiao, T.; Yu, K.; Deng, Y.; Peng, S. Boron Nitride Nanotubes Reinforce Tricalcium Phosphate Scaffolds and Promote the Osteogenic Differentiation of Mesenchymal Stem Cells. J. Biomed. Nanotechnol. 2016, 12, 934–947.
  136. Li, R.; Lin, J.; Fang, Y.; Yu, C.; Zhang, J.; Xue, Y.; Liu, Z.; Zhang, J.; Tang, C.; Huang, Y. Porous Boron Nitride Nanofibers/PVA Hydrogels with Improved Mechanical Property and Thermal Stability. Ceram. Int. 2018, 44, 22439–22444.
  137. Belaid, H.; Nagarajan, S.; Barou, C.; Huon, V.; Bares, J.; Balme, S.; Miele, P.; Cornu, D.; Cavailles, V.; Teyssier, C.; et al. Boron Nitride Based Nanobiocomposites: Design by 3D Printing for Bone Tissue Engineering. ACS Appl. Bio Mater. 2020, 3, 1865–1874.
  138. Aki, D.; Ulag, S.; Unal, S.; Sengor, M.; Ekren, N.; Lin, C.-C.; Yilmazer, H.; Ustundag, C.B.; Kalaskar, D.M.; Gunduz, O. 3D Printing of PVA/Hexagonal Boron Nitride/Bacterial Cellulose Composite Scaffolds for Bone Tissue Engineering. Mater. Des. 2020, 196, 109094.
  139. Özmeriç, A.; Tanoğlu, O.; Ocak, M.; Çelik, H.H.; Fırat, A.; Kaymaz, F.F.; Koca, G.; Şenes, M.; Alemdaroğlu, K.B.; Iltar, S.; et al. Intramedullary Implants Coated with Cubic Boron Nitride Enhance Bone Fracture Healing in a Rat Model. J. Trace Elem. Med. Biol. 2020, 62, 126599.
  140. Baghdadi, I.; Zaazou, A.; Tarboush, A.B.; Zakhour, M.; Özcan, M.; Salameh, Z. Physiochemical Properties of a Bioceramic-Based Root Canal Sealer Reinforced with Multi-Walled Carbon Nanotubes, Titanium Carbide and Boron Nitride Biomaterials. J. Mech. Behav. Biomed. Mater. 2020, 110, 103892.
More
This entry is offline, you can click here to edit this entry!
Video Production Service