Next Article in Journal
Modified Mycotoxins, a Still Unresolved Issue
Previous Article in Journal
Quantitative Measurements of Pharmacological and Toxicological Activity of Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sol–Gel Routes toward Ceramic Nanofibers for High-Performance Thermal Management

1
School of Textile and Clothing, Nantong University, Nantong 226019, China
2
Key Laboratory of Multifunctional Nanomaterials and Smart Systems, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou 215123, China
3
School of Civil Aviation, Northwestern Polytechnical University, Xi’an 710072, China
4
National Equipment New Materials & Technology (Jiangsu) Co., Ltd., Suzhou 215101, China
*
Authors to whom correspondence should be addressed.
Chemistry 2022, 4(4), 1475-1497; https://doi.org/10.3390/chemistry4040098
Submission received: 7 October 2022 / Revised: 28 October 2022 / Accepted: 2 November 2022 / Published: 8 November 2022
(This article belongs to the Section Chemistry at the Nanoscale)

Abstract

:
Ceramic-based nanofiber materials for high-performance thermal management have drawn increasing attention owing to their high-temperature resistance, efficient thermal insulation, superior mechanical flexibility, as well as excellent physical–chemical stability. We present an overview of the ceramic-based nanofiber obtained by sol–gel routes for high-performance thermal management, including the materials, the fabrication methods of the sol–gel route, and their application for thermal management. We first provide a brief introduction to the ceramic-based nanofibers. The materials and fabrication methods of the sol–gel route are further discussed in the second part, including the kinds of nanofibers such as oxide, carbide, and nitride, and the methods such as centrifugal spinning, electrospinning, solution blow spinning, and self-assembly. Finally, their application for thermal management is further illustrated. This review will provide some necessary suggestions to researchers for the investigation of ceramic-based nanofibers produced with the sol–gel route for thermal management.

1. Introduction

Ceramic-based fiber materials, especially the micron–nano-level ceramic fibers, have drawn increasing attention in recent years. Efforts toward developing ceramic-based fibers have been increasing for nearly a decade, dominated by research and application reports from China, accounting for 40% of the total publications among all countries [1].
The most common ceramic fibers are usually fabricated by the slurry method or molten liquid spinning, both methods utilizing a ceramic slurry made of a micron–nano-level ceramic particle as the precursor, the inherent properties of which, such as great hardness and brittleness of ceramic materials, lead to an insufficient toughness of the ceramic-based fiber materials, and seriously limit their applications. On the other hand, flexible ceramic fiber products have been in huge demand for high-performance thermal management in high temperatures, owing to their superior high-temperature resistance, great mechanical flexibility, and excellent chemical stability.
Researchers, therefore, have applied significant effort towards the preparation of flexible ceramic-based nanofibers, especially for two- or three-dimensional products, such as ceramic-based nanofiber films or aerogels, because they usually possess ultra-low density, high porosity, large specific surface area, high-temperature resistance, oxide resistance, and low thermal conductivity in wide temperature ranges. For the preparation of ceramic-based nanofibers, there are several methods: the slurry method, the chemical vapor deposition method, the polymer precursor conversion, the molten liquid spinning, and the sol–gel method, among which the sol–gel route method dominates due to its low cost, low operating temperature, and excellent production efficiency. In addition, ceramic-based nanofibers have been widely applied due to their superior properties in the industries of gas filtration [2], sound absorption [3,4], thermal insulation [5,6,7], luminescence [8], microwave absorption [9], electromagnetic shielding [10], and others.
In this review, we summarized flexible ceramic-based nanofibers for the purpose of thermal management, including the preparation method, the materials, the thermal insulation mechanisms, and their applications. We first introduce the sol–gel routes of the ceramic-based nanofibers. Next, we briefly introduce the materials utilized in the fabrication of the ceramic-based nanofibers, including the oxides, carbides, and nitrides. Finally, applications of ceramic-based nanofibers in thermal management are discussed.

2. Nanofibers Prepared by Sol–Gel Routes

Nanofibers can be divided into two types: organic and inorganic, according to their chemical components. The organic nanofibers include biofibers such as cellulose-based fibers and protein fibers, and chemical fibers such as polyethylene (PE), polypropylene (PP), polyvinyl acetate (PVA), and polyvinyl chloride (PVC). Inorganic nanofibers include oxide fiber, carbide fibers, nitride fibers, and others. Owing to organic fibers’ inherent lack of high-temperature resistance, researchers aiming at developing nanofibers for high-temperature applications mainly focus on low-cost fabrication of the inorganic fibers, especially those which could be used in high-temperature conditions, such as the SiC [11], BN [12], SiO2 [5], Al2O3 [13], and ZrO2 [6] fibers. Among them, the sol–gel process provides an effective approach for ceramic-based nanofiber fabrication.

2.1. Fabrication Routes

2.1.1. Centrifugal Spinning

Centrifugal spinning is a method for the preparation of nanofibers that utilizes the centrifugal force generated by a high-speed rotation to convert the liquid into fibers. A schematic of typical centrifugal spinning equipment is displayed in Figure 1, which was employed to fabricate the fibers. The liquid is usually added to the container with a spinneret. For molten ceramic raw fibers, a high-temperature alloys container is usually used, maintaining its temperature over 1500 °C with a very high rotational speed, usually higher than 3000 rpm, to keep the raw ceramic material in a molten state [14]. When the centrifugal stress exceeds the surface tension of the liquid, the liquid will be injected from the spinneret to form several jets. The liquid flowing out of the sidewall will be stretched by the centrifugal force and air frictional force and reformed into fibers with a diameter of 1 μm or higher. To obtain nanofibers with an average diameter below 1 μm, liquid precursor centrifugal spinning is always employed as one of the most popular approaches for manufacturing nanofibers, such asZrO2 fibers [15], SiO2 nanofibers [16], and Al2O3 nanofibers [17], etc. The schematic diagram of centrifugal spinning and corresponding equipment is shown in Figure 1.

2.1.2. Electrospinning

Besides centrifugal spinning, electrospinning is also a widely used method utilized for the fabrication of nanofibers [19,20]. Figure 2 and Figure 3 show a schematic of an electrospinning process for manufacturing nanofibers [21,22]. Usually, a high-voltage electrical field is built between the nozzle and collector. Nanofibers formed at the nozzle owing to a much higher electrostatic force than the liquid jet’s surface tension. During the spinning, nanofibers are formed after the evaporation of the solvents. The ultrafine nanofibers are finally obtained on the collector, including organic nanofibers and Al2O3 nanofibers. Electrospinning is a typical method for nanofiber preparation, and it contains at least three typical steps: (1) preparation of a spinning solution consisting of precursors; (2) solution spinning under an electric field; and (3) calcination at high temperatures to obtain the inorganic nanofiber. Therefore, the advantages of electrospinning include structure and morphology control simply by adjusting the spinning parameters such as the composition of the spinning solution, electric field, and calcination conditions.

2.1.3. Solution Blow Spinning

Solution blow spinning is also one of the most effective fiber preparation approaches. It can produce a series of ceramic and polymer fibers with high efficiency and low cost [23,24]. Figure 4 displays a schematic diagram of a typical solution blow spinning, including a compressed air source, a collector, and a solution conveying device. During solution blow spinning, a solution is extruded through the inner nozzle and a high-speed gas flow is ejected through the outer nozzle, in which the solution forms a jet under the shear action of high-speed gas flow, which is further split and refined to form fibers. Similar to the electrospinning method, solution blow spinning also starts with the preparation of a spinning solution and ends with calcination, where the precursor firstly forms the raw fibers by blow spinning and finally calcination at high temperatures to form the ceramic fibers.
Solution blow spinning is a simple procedure compared to electrospinning as shown in Figure 5, which makes it popular for the fabrication of nanofibers [25,26,27,28]. Compared to electrospinning, the solution blow spinning is high-voltage-free and it works with the high-speed gas flow as the driving force, thus it draws more attention for possessing a higher spinning efficiency.

2.1.4. Self-Assembly (Freeze-Drying)

The diameters of nanofibers are usually below 200 nm, and the lengths are several microns, which makes it hard to form the applicable materials. Therefore, self-assembly is an approach for combining nanofibers to build a novel morphology with adjustable structures. The mechanism of nanofiber self-assembly is the formation of hydrogels or a new nanofibric microstructure, which contains two routes: the liquid phase method and the solid phase method. The liquid phase method employs the liquid as a medium to dry a mixture of fiber precursor and liquid by removing the liquid phase to obtain hydrogels, such as the BN aerogel preparation process by Li et al. [30], which shows a preparation process of BN nanoribbon aerogels by the self-assembly method. In addition, Si et al. [5] introduced a SiO2 nanofiber aerogel shown in Figure 6, which was also fabricated by the self-assembly method. The SiO2 nanofiber aerogel showed robust mechanical performance, such as twisting and excellent flexibility.

2.1.5. Other Routes

Except for the routes mentioned above, there are some other routes with the sol–gel method for the production of the nanofibers, such as the polymer conversion method, which is a method for preparing high-temperature resistant nanofibers from Si-, N-contained polymer, etc. [31]. Clearly, it can be seen that there are differences among the above-mentioned method. The differences are summarized in Table 1 by Jia [1].

2.2. Nanofibers

2.2.1. Oxide Nanofibers

Oxide nanofibers possess the qualities of high-temperature resistance, oxidation resistance, high mechanical strength, great electrical, thermal, and acoustic insulation, as well as chemical stability. Among them, SiO2, Al2O3, Mullite, ZrO2, ZnO, and TiO2 are usually employed to fabricate the corresponding ceramic-based nanofibers.
(1)
SiO2 nanofibers
As the most common oxide fibers, SiO2 fibers are widely utilized in fields such as terrarium, electromagnetic transmission, and thermal insulation. The SiO2 nanofiber is mainly employed as efficient thermal insulation at high temperatures. Therefore, most researchers are focusing on improving the thermal insulation performance of the SiO2 nanofibers in a wide temperature range.
Si et al. [5] fabricated a SiO2 nanofiber with an average diameter of 206 nm by a sol–gel electrospinning method. SiO2 nanofiber aerogel with a mixed structure of layer-nanofibers-layer was further prepared by a fibrous freeze-shaping method using the flexible SiO2 nanofibers as building blocks. Wang et al. [32] prepared flexible SiO2 nanofibers with an average diameter of 162 nm by the sol–gel electrospinning method. The flexible SiO2 nanofibers were used to construct a super elasticity biomimetic SiO2 nanofibrous aerogel. Wang et al. [33] further obtained a ZrO2/SiO2 fiber membrane possessing a prominent removal efficiency of 85% and excellent adsorption amount of 43.8 mg·P·g−1 in 30 min toward phosphates by modifying SiO2 nanofibers with ZrO2. Mi et al. [34] obtained 3D SiO2 nanofibers with a low bulk density of 16 mg·cm−3 and a high specific surface area of 6.5 m2·g−1 by the sol–gel self-assembly and electrospinning method. The deposition and growth behavior of the 3D fibers were studied. The results indicated that the electrostatic repulsion among fibers caused a loose packing of fibers, and the rapid solidification of the fibers contributed to maintaining the 3D structure. Mao et al. [35] prepared SiO2 nanofiber membranes by the combination of electrospinning and sol–gel methods, which possessed excellent flexibility of 0.0156 gf·cm−1, a tensile strength of 5.5 MPa, a thermal stability up to 1000 °C, and a high infiltration efficiency. Huang et al. [36] developed SiO2 nanofibers by a combination of electrospinning and sol–gel methods, which possessed an ultralow density of 7.9~37 mg·cm−3, excellent resilient compressibility of 0.75 MPa at −85% strain, and a corresponding elastic modulus of 0.72 MPa. Yan et al. [21] employed sol–gel electrospinning to synthesize SiO2 nanofibers films with a silk-like softness of <31 mN, a low density of <0.36 g·cm−3, and robust fire resistance of 1000 °C, which displayed large electrolyte uptakes of >900% and high thermal insulation performance, enhancing the rate capability and safety of lithium batteries. Lu et al. [29] utilized a high-pressure gas spinning method to fabricate nanofibers/nanotubes with an individual setup of a throughput of 10 g·s−1. In addition, a mechanism for the extrusion method was proposed based on flow mechanics and the experimental results, as shown in Figure 7.
Ren et al. [37] prepared SiO2 nanotube fibers with wall thickness < 100 nm and outer diameters in the range of 300~500 nm by centrifugal jet spinning. By using the same technology, Ren et al. [38] further fabricated a series of SiO2 micro-/nanofibers with hollow or porous internal structure, whose solid cross sections were 364 and 781 nm, and the hollow cross sections were 458 and 216 nm, respectively. Tepekiran et al. [16] produced SiO2 nanofibers via centrifugal spinning and subsequent calcination. The SiO2 nanofibers fabricated by calcinated the 15 wt.% TEOS/PVP sample showed the highest filtration performance, and the average fiber diameter was the lowest, around 521 ± 308 nm. The nanofibers showed an enhanced filtration efficiency of around 75.89%.
(2)
Al2O3 nanofibers
Similar to the SiO2 fibers, the Al2O3 fibers are also widely applied in the fields of electromagnetic transmission and thermal insulation, especially for thermal insulation at high temperatures over 1000 °C. The Al2O3 nanofibers are developed to solve the problems of growth in thickness and increase in thermal conductivity during utilization.
Mahapatra et al. [13] synthesized ultra-fine α-Al2O3 nanofibers by using a mixture of PVP/ethanol solution and aluminum acetate sol via the electrospinning technique. The results showed that pure and crystalline α-Al2O3 nanofibers with a diameter range from 100~500 nm and BET surface area of 40 m2·g−1 were formed. Wang et al. [39] fabricated mesoporous alumina fibers via an electrospinning technique combined with a sol–gel method, which exhibited uniform mesopores with a diameter range from 130~200 nm and a BET specific surface area up to 264.1 m2·g−1. By transferring the γ-Al2O3 to α-Al2O3 at 1000 °C, the mesopores disappeared with a grain growth. The γ-Al2O3 showed a great adsorption to Congo red with a maximum adsorption capacity of 781.25 mg·g−1. Wang et al. [40] further fabricated γ-Al2O3 nanofiber membranes by the electrospinning method. The nanofibers possessed a high aspect ratio, small diameter of 230 nm, high tensile strength of 2.98 MPa, and thermal stability of ~900 °C. Furthermore, the membrane exhibited great filtration performance for dioctyl phthalate, with a filtration efficiency of 99.8% at a pressure drop of 239.12 Pa when the basis weight was 9.28 g·m−2. Filtration efficiency over 99.97% could be achieved when the basis weight was 11.36 g·m−2. Song et al. [41] prepared Al2O3 nanofibers by the sol–gel electrospinning method with different polymer templates, such as polyvinyl alcohol, polyvinyl butyral, and polyvinyl pyrrolidone. Three kinds ofγ-Al2O3 nanofibers were obtained and they exhibited completely different properties such as continuous structure, densification degrees, crystallization temperatures, grain sizes, tensile strength, and elastic modulus. Li et al. [42] obtained flexible Al2O3 nanofiber mats by combing the sol–gel method with the solution-blowing process and a subsequent thermal treatment process. The as-spun nanofibers were formed after the solvent evaporation and drawing of the gas flow, and converted into γ- or α-Al2O3 under a treatment in air at 900~1000 °C. The diameter range 2~5 μm and influenced by the spinning parameters, such as gel viscosity and high-pressure gas flow. Sedaghat et al. [17] manufactured an Al2O3 nanofibers mat by a sol–gel spinning method using a self-laboratory designed centrifugal spinneret. It was found that θ-Al2O3 was the main phase below 800 °C, which was transformed to α-Al2O3 at 1000~1200 °C with an optimum silica percent of 4 wt.%. The met was constructed by the nanofibers constituting a network with pore diameter ranges from 100 nm to 10 μm. When the silica content was lower than 10 wt.%, the fiber diameter had no effect. However, the grain size decreased from 200 nm to 100 nm or lower when the silica content was increased. Akia et al. [43] fabricated aluminum isopropoxide and PE fine composite fibers utilizing the response surface methodology. The γ- and α-Al2O3 nanofibers showed a diameter of 304 nm, while the γ-Al2O3 nanofibers possessed a mesoporous structure with a BET surface area of 261 m2·g−1.
(3)
Mullite nanofibers
Mullite is a stable phase composed the crystal of SiO2 and Al2O3 with an Al2O3 content of 72~78 wt.%, which possesses better coarsening resistance as compared to the single component of SiO2 or Al2O3. Therefore, the mullite nanofibers are also widely applied as alternatives to SiO2 and Al2O3, especially for thermal insulation for long-time service at high temperatures.
Liu et al. [44] prepared mullite-based nanofibrous aerogels via the gel-casting and freeze-drying methods with a series of different mullite of SiO2:Al2O3 = 3:2~0, as shown in Figure 8. All the mullite nanofibrous possessed minor pores on the surface. They further overlapped to form nanofiber aerogel with multilevel porous structure, ultralow density of 34.64~48.89 mg·cm−3, low thermal conductivity of 0.03274~0.04317 W·m−1·K−1, and great physical and mechanical properties. Zadeh et al. [45] synthesized 3Al2O3-2SiO2 nanofibers by combining the sol–gel and electrospinning methods with a mixture sol of aluminum isopropoxide, hydrated aluminum nitrate, and tetraethylorthosilicate. The optimal amount of PVB in the polymeric electrospinning solutions ranged from 4~6 wt.%, and the obtained mullite nanofibers were pure, smooth, and uniform, with diameter of 85~130 nm. Song et al. [46] prepared flexible mullite nanofibers by an electrospinning method through the conventional diphasic mullite sol–gel route. The precursor sol was prepared by mixing aluminum acetate, colloidal silica, and the polyvinyl pyrrolidone. The mullite nanofiber was obtained after calcining at 1000 °C with an optimum Al/Si molar ratio of 2.98 and an elastic modulus of 25.18 ± 1.29 GPa. Xian et al. [47] fabricated mullite nanofibrous aerogels via the freeze-casting method by controlling the growth of ice crystals to adjust microstructure. The aerogel with high agarose content exhibited a relatively high compressive strength of 252.41 kPa owing to a high density and uniform porous structure. Song et al. [22] prepared continuous mullite nanofibers by the conjugate electrospinning technique combined with the sol–gel method by using the aluminum acetate stabilized with boric acid and tetraethyl orthosilicate. The obtained nanofibers showed an elastic modulus of 12.27 ± 1.77 GPa, and a tensile strength of 32.21 ± 3.73 MPa for the aligned mullite nanofiber bundle was obtained. Costa et al. [48] produced submicrometric mullite fibers via the solution blow spinning method. The mullite fibers were crack-free, possessed a diameter of 800 nm, and can be fabricated into fibrous mats showing a high surface area of 24 m2·g−1.
(4)
ZrO2 nanofibers
Ruiz et al. [49] used electrospinning for the preparation of PVP–zirconium acetate nanofibers in non-woven cloths. The ZrO2 nanofibers showed a high aspect ratio and a thin thickness of 200 nm. Chen et al. [50] prepared ZrO2 nanofiber membranes by electrospinning combined with the sol–gel methods. By incorporating yttrium oxide, the as-prepared ZrO2 nanofiber membranes can be changed from extremely fragile to robust and offer greater flexibility for further application. Wang et al. [51] produced a resilient and heat-resisting yttria-stabilized ZrO2 nanofiber sponge by a scalable solution blow spinning process. The porous 3D sponge showed a low density of 20 mg·cm−3, and 99.4% filtration efficiency for aerosol particles (20~600 nm) with a low-pressure drop of only 57 Pa when the airflow velocity was 4.8 cm·s−1. In addition, the sponge maintained a high filtration efficiency of 99.97% for PM 0.3~2.5 in an airflow velocity of 10 cm·s−1. Wang et al. [6] fabricated a superfine ZrO2 nanofiber ceramic aerogel with densities varying from 8 to 40 mg·cm−3, which exhibited a maximum of 29.6 mJ·cm−3 in energy density at 50% strain at a compression of 20%. The aerogel exhibited excellent resilience with residual strains of only ~1% at 800 °C after 10 cycles of 10% compression strain and good recoverability after compression at ~1300 °C.
(5)
Other oxide nanofibers
SiO2, Al2O3, and mullite are three of the main materials that are focused on the field of thermal insulation. There are other oxide nanofibers for other applications, such as nanofibers of ZnO [52,53,54,55,56], TiO2 [57,58,59,60,61,62,63,64,65], CeO2 [66], and CuO [67], etc.

2.2.2. Carbide Nanofiber

Carbide nanofibers are also widely employed as materials for thermal insulation, such as SiC, SiCN, and ZrC. Among them, SiC is the most investigated nanofiber. Owing to the covalent Si-C bond in SiC, SiC nanofiber displays excellent chemical stability and high mechanical resistivity below 1500 °C. Moreover, the performance of oxide high-temperature resistance could be further improved by doping or coating.
Su et al. [11] fabricated SiC nanowire aerogels consisting of a large number of interwoven 3C-SiC nanowires by chemical vapor deposition, which possessed an ultra-low density of 5 mg·cm−3, a large strain recoverable compressibility over 70%, a low thermal conductivity of 0.026 W·m−1·K−1, and high organic solvent absorption properties of 130~237 g·g−1. An et al. [68] synthesized a SiC nanofiber aerogel by rapidly stirring, freeze-drying, and applying a high-temperature treatment based on the fine SiC fibers prepared by electrospinning. The aerogel exhibited a low density of 0.039~0.041 g·cm−3, and a low thermal conductivity of 0.025~0.031 W·m−1·K−1 for excellent thermal insulation. Liu et al. [69] employed the PMMA as a sacrificial template and co-axial electrostatic spinning to prepare N-doped hollow SiC fiber mats, as shown in Figure 9. The mats possessed a cavity wall thickness of approximately 1.5 μm, a low density of 0.218 g·cm−3, and low thermal conductivity of 0.039 W·m−1·K−1. Li et al. [70] prepared micro-scale SiC nanofiber mats with great thermal stability and semi-conductivity by blowing off an oil-in-water (O/W) precursor, which was followed by an emulsification process, thermal curing, and high-temperature calcination.
Wang et al. [71] used electrospinning technology to obtain ultrafine ZrO2/SiC fibers. By pyrolysis at different temperatures, the fibers possessed a radial gradient composition and high-temperature stability over 1800 ℃, as shown in Figure 10. Wu et al. [72] obtained a hydrophobic SiOC nanofiber membrane without any modification by the pyrolysis reaction of electrospun polycarbosilane nanofibers, as displayed in Figure 11. The nanofibers exhibited consistent hydrophobicity and strong mechanical properties throughout the pH and high temperature, as displayed in Figure 4. To prepare SiOC fibers with homogeneous morphology, the electrostatic spinning of a solution containing two (MK and H44) methyl silicones was employed. The effect of the processing process on the fibers was investigated by Guo et al. [73]. The results indicated that the introduction of 20 vol% N, N-dimethylformamide (DMF) was able to decrease the diameter of the as-spun fibers from 2.72 ± 0.12 μm to 1.65 ± 0.09 μm. For the H44/DMF system, the diameter of the spun fiber obtained by adding 50 vol% chloroform was 1.61 ± 0.16 μm. Both the pyrolyzed MK and H44-derived SiOC fibers have a uniform, defect-free morphology, small average diameter, and narrow size distribution.
Dong et al. [74] prepared a SiOC nanofiber aerogel via electrospinning assembly by the gel-casting and freeze-drying method, as exhibited in Figure 10. The aerogel possessed a low density of 46.87~128.48 mg·cm−3, low thermal conductivity of 0.0302~0.0440 W·m−1·K−1, high compressive strength of 18~167 kPa, and high specific surface area of 12.48~34.87 m2·g−1. SiCN nanowires with efficient electromagnetic wave absorption properties were prepared by electrostatic spinning and high-temperature annealing (in a nitrogen atmosphere) by Wang et al. [75], which exhibited ultra-flexibility under winding and bending. The nanowires had an optimal reflection loss (RL) of −53.1 dB, and an effective absorption broadband (EAB) of 5.6 GHz.

2.2.3. Nitride Nanofiber

(1)
BN nanofiber
Similar to carbon, BN possesses a hexagonal, cubic, or tubular structure, and exhibits high thermal oxidation resistance and high radiation absorption capacity owing to its strong structural atomic connections of the B-N bond. Moreover, BN exhibits high thermal conductivity and desired mechanical property, enabling their potential utilization as thermal mechanical structure.
Song et al. [76] manufactured a BN aerogel by a template-assisted approach of low-pressure chemical vapor deposition on graphene–carbon nanotubes at 900 °C. It possessed an ultralow density of 0.6 mg·cm−3, an ultrahigh specific surface area up to 1051 m2·g−1, and high hydrophobicity, which resulted in excellent oil absorption of up to 160 times by weight. Xue et al. [12] prepared h-BN nanofibers by employing the B2O3 in situ chemical vapor deposition on 3D-SiO2/N-doped tubular graphitic cellular foams. The foam was built by the h-BN nanofibers and possessed an interconnective nanotubular architecture. The foam featured with ultralight weight, thermal stability, high porosity of 98.5%, remarkable shape recovery for cycling compressed with 90% deformations, and excellent high-capacity adsorption–separation for oil/water systems. Li et al. [30] built a BN nanofiber aerogel with excellent temperature-invariant super-flexibility by a melamine diborane precursor. As exhibited in Figure 12, the aerogel was assembled by hydrogen bonding and possessed outstanding compressing/bending/twisting elasticity, cutting resistance, recoverable properties, and excellent mechanical super-flexibility over a wide temperature range of −196~1000 °C.
(2)
Si3N4 nanofiber
Si3N4 is another nitride ceramic that is widely investigated for thermal insulation and electromagnetic wave shielding, etc. Su et al. [77] produced ultralight α-Si3N4 nanobelt aerogels with density ranges from 1.8~9.6 mg·cm−3. The nanobelt aerogels owned resilient compressibility of a recoverable strain of 40~80%, great fire resistance at 1200 °C, excellent thermal insulation with a low thermal conductivity of 0.029 W·m−1·K−1, electronic wave transparency of a dielectric constant of 1~1.04, and a dielectric loss of 0.001~0.004. Huo et al. [78] fabricated 3D Si3N4 nanofiber-knitted ceramic foams via in situ reactive synthesis derived from silicon foams. The foams consisted of 3D nanofibers with a diameter of 15~100 nm. Zhang et al. [79] prepared Si3N4 nanowires via the vapor–solid process,. The diameters of the nanowires ranged from 20~40 nm, and they had a specific surface area of 50.47 m2·g−1, showing their promising applications in filtration, thermal insulation, and catalyst supports.
(3)
Other nitride nanofibers
Except for the BN and Si3N4, some other important nitrides are valued for their excellent mechanical property and chemical stability, such as the GaN [80] used for the third semiconductor and AlN.

3. Materials Prepared by the Nanofibers and Their Thermal Management Applications

Ceramic-based nanofiber materials draw significant attention for thermal insulation applications owing to their low density, low thermal conductivity, high temperature resistance, good chemical stability, and good flexibility. Therefore, research on its application, especially at high temperatures, has been the focus in recent years.
To achieve high elasticity, Si et al. [5] prepared ceramic nanofiber aerogels (CNFAs) with superelastic layered structures by combining silica nanofibers with an aluminum-borosilicate matrix. Owing to their ultra-low density, fast recovery at 80% strain, zero Poisson’s ratio, and temperature invariant super-relaxation at 1100 °C, the CNFAs have been used in insulation, catalysis, adsorption, energy, and acoustics, etc. They open up a wide range of technological implications.
The application of ultralight and porous nanostructures has been considerably limited by their brittleness. Wang et al. [6] prepared ceramic sponges consisting of a large number of twisted ceramic nanofibers based on oxide ceramics by a cost-effective blow-spinning technique. These ceramic sponges exhibited ultra-low density, high elasticity, and energy absorption, providing new insights into the design of porous cell structures for high-temperature applications.
Jia et al. [3] developed an anisotropic layered SiO2-Al2O3 composite ceramic (SAC) sponge using a facile sol–gel solution blow spinning technique, which successfully solved the problem of the complicated preparation process of ceramic sponge materials. This sponge exhibited temperature invariant compression elasticity from −196~1000 ℃, good thermal conductivity of 0.034 W·m−1·K−1, and a density as low as 10 mg·cm−3. The excellent properties are promising in the field of heat insulation and sound absorption.
Su et al. [11] reported highly porous 3D silicon carbide nanowire aerogels (NWAs) as shown in Figure 13. The NWAs had ultra-low density (5 mg·cm−3), high porosity (99.8%), and good heat resistance properties (0.026 W·m−1·K−1 at room temperature in N2), etc. These excellent properties make the NWAs promising in the field of fireproof materials and high-temperature insulation.
The creation of 3D nano aerogels (NFAs) is challenging. Si et al. [4] combined electrostatic spinning techniques and fiber freezing to prepare superelastic, layered cellular structured NFAs. The method enables the assembly of essentially layer-deposited electrostatic spun nanofibers into elastic aerogels with tunable density and shape. The NFAs exhibited a density of >0.12 mg·cm−3. The bionic silica nanofiber (SNF) aerogel synthesized by Wang et al. [32] exhibited ultra-low density (>0.25 mg·cm−3) and temperature invariant superelasticity up to 1100 ℃.
Liu et al. [78] used co-axial electrostatic spinning to prepare nitrogen-doped hollow silicon carbide fiber mats, in which PMMA was used as a sacrificial template to stabilize and pyrolyze the PCS/PMMA precursor. The cavity wall thickness of the hollow fiber was approximately 1.5 μm, while the mats possessed low density (0.218 g·cm−3), high thermal conductivity (0.039 W·m−1·K−1), good flexibility and thermal stability. The properties of the mats suggest promising application as high-temperature thermal insulators. Dou et al. [81] synthesized a layered cellular structured silica nanofiber aerogel with adjustable density below 10 mg·cm−3, excellent thermal insulation property with thermal conductivity of 0.029 W·m−1·K−1, and ultra-low dielectric constant due to the high porosity of NBAs. The aerogels showed excellent refractoriness and high-temperature resistance, as exhibited in Figure 14.
Su et al. [82] proposed the use of directional freeze casting and heat treatment to prepare SiC@SiO2 nanowire aerogels with macroscopic pores to reduce the insulation (as exhibited in Figure 15). These aerogels have ultra-low thermal conductivity 14 mW·m−1·K−1 and excellent thermochemical stability (even at 1200 °C under butane blow gun), which are rational thermal insulation materials under extreme conditions. The new generation of elastomeric ceramic aerogels (CAGs) is distinguished in the field of thermal insulation by solving the problem of extreme conditions, where insulation materials need to be resistant to sharp thermal shocks and prolonged exposure to high temperatures [83]. In addition, Hu et al. [84] summarized the present researched thermal insulation materials utilized over 800 °C, in which the high-efficiency thermal insulation is almost built by the sol–gel routes as nanoparticle aerogels and nanofibers aerogels. They showed excellent thermal insulation performance with low thermal conductivities, as listed in Table 2. As has been mentioned above, ceramic-based nanofiber materials exhibit an excellent thermal insulation effect, especially at a relatively wide temperature range from negative up to 1000 °C, which indicates a direction to the development the more efficient thermal insulation materials applied in the conditions with higher temperature over 1500 °C, wider temperature range, or strong radiation environments.
High-temperature insulation is widely used in various fields, such as aerospace, construction, and other fields.
Wang [85] studied a sapphire optical fiber high-temperature sensor sealed in an alumina ceramic tube–ceramic sleeve structure. The sensor has a measurable range of 25–1550 °C, good stability, high-temperature detection accuracy, and a very broad application prospect in high-temperature fields such as aircraft engines and gas turbines. Solid oxide fuel cells (SOFCs) are the most flexible, cleanest, and most efficient chemical–electrical energy conversion systems among all types of fuel cells. Chen [86] avoided the disadvantages of the fuel by surface modification of electrocatalytic activity consisting of conformal PNM films and dissolved PrOx nanoparticles, which enhanced hybrid catalyst coatings. He [87] explored a new method of membrane-based propane POX, in which propane reacts with permeate oxygen to syngas in the presence of the Ru-Ni catalyst (Figure 16). The membrane reactors are superior to conventional reactors in terms of safety and syngas concentration, so they have great application prospects as pre-reformers for solid oxide fuel.
Gao [88] provides a perspective on solid oxide fuel cells (SOFCs) operating at low temperatures (LT) (400~650 °C). The LT-SOFCs exhibited good battery performance at low temperatures, offer potential advantages in traditional SOFC applications, and may be suitable for new portable and transport power applications. Higher-temperature SOFCs are currently being commercialized. Yang [89] focused on the research of lightweight layered silicon pore gradient ceramic aerogel foam. The high porosity and large pore gradient of the aerogel endowed it with excellent thermal properties. Meanwhile, the hydrophobic PGA foams have mechanical stability and fire resistance in humid environments, as shown in Figure 17, which plays an important role in building materials.
Cheng [90] prepared a new lightweight needle-like carbon fiber felt/phenolic resin (NCF-PR) aerogel composite. The aerogel composite had a unique bird’s nest structure, which provides excellent thermal insulation and ablation performance in the arc jet wind tunnel simulation environment, as exhibited in Figure 18. They showed great application potential in aerospace thermal insulation.
Jin [91] developed a clay aerogel with both hydrophobic and mechanical robustness. The aerogel exhibited ultra-low density, limiting oxygen index of up to 90%, and fire resistance at 1000 °C for ten minutes. Thus, it is a very promising material for construction and aerospace owing to its heat insulation and fire resistance. Nguyen [92] reported a family of triisocyanate aerogels formed by crosslinking amine-capped polyimide oligomers and triisocyanates. The backbone chemistry, chain length, and polymer concentration determine its low density, large surface area range, and high compression modulus. Due to its low cost, the aerogel can be widely used. In the case that its low temperature stability could be solved, the aerogel would attract more attention in aerospace application. Compared to traditional thermal management materials, the materials obtained from the sol−gel route possesses higher strength and stiffness by selecting appropriate silane precursors and polymer reinforcement. Randall [93] discussed some strategies to improve the strength and elasticity of silica aerogels in recent years, as listed in Figure 19, so that silica aerogels can be more suitable for future applications.

4. Conclusions

Developments and applications of ceramic-based nanofiber materials in the last decade have been summarized in this review. The types of nanofibers including oxide, carbide, nitride, and other ceramic-based nanofibers, and their preparation methods, such as centrifugal spinning, electrospinning, solution blow spinning, and self-assembly, which have been employed for realizing the superfine nanofibers, are summarized. The development of fabrication methods has prompted the applications of the ceramic-based nanofibers in the fields of sound insulation, electromagnetic wave insulation, and thermal insulation. In particular, the ceramic-based nanofibers find specific applications in high-temperature thermal insulation. Based on the above summarization, it can be concluded that the production of superfine ceramic-based nanofibers is the development direction for the high-performance thermal management materials, and the sol−gel route is the most efficient approach. There still exist some problems that will prevent it from rapid development and commercialization. Some tips to avoid or solve these problems for the development and application of the ceramic-based nanofibers are proposed:
(1)
Avoid the utilization of polymers to enhance the quality of nanofibers. Polymers result in fiber diameter reduction, formation of porous surface, and decrease in mechanical capacity. The way to eliminate this phenomenon is to enhance the solid content as much as possible and avoid polymer utilization.
(2)
Multiphase component nanofibers. For some single-component, ceramic-based nanofibers, their performance will decrease because of the nanometer effect. It is therefore necessary to obtain multiphase component nanofibers by adding some other components to improve their performance, such as the YSZ-Al2O3 nanofibers.
(3)
Improvement for efficient commercialization. The solution blow spinning possesses the highest efficiency to prepare nanofiber-based products, such as films and sponges, which ought to receive more research attention.

Author Contributions

J.Z. (Jing Zhang), J.Z. (Junxiong Zhang), Q.S., X.Y., X.M., and J.W. co-write the manuscript, J.Z. (Junxiong Zhang) and J.W. supervised and obtained funding for the project. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Talent Introduction Project Foundation of Nantong University (135421615077); Large Instruments Open Foundation of Nantong University (KFJN2237); the National Natural Science Foundation of China (91963124), the Suzhou Municipal Science and Technology Bureau (SJC2021008); China Postdoctoral Science Foundation (2021M702665); and Natural Science Foundation of Shaanxi Province (2022JQ-482).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jia, C.; Xu, Z.; Luo, D.; Xiang, H.; Zhu, M. Flexible Ceramic Fibers: Recent Development in Preparation and Application. Adv. Fiber Mater. 2022, 4, 573–603. [Google Scholar] [CrossRef] [PubMed]
  2. Jia, C.; Liu, Y.; Li, L.; Song, J.; Wang, H.; Liu, Z.; Li, Z.; Li, B.; Fang, M.; Wu, H. A foldable all-ceramic air filter paper with high efficiency and high-temperature resistance. Nano lett. 2020, 20, 4993–5000. [Google Scholar] [CrossRef] [PubMed]
  3. Jia, C.; Li, L.; Liu, Y.; Fang, B.; Ding, H.; Song, J.; Liu, Y.; Xiang, K.; Lin, S.; Li, Z.; et al. Highly compressible and anisotropic lamellar ceramic sponges with superior thermal insulation and acoustic absorption performances. Nat. Commun. 2020, 11, 3732. [Google Scholar] [CrossRef] [PubMed]
  4. Si, Y.; Yu, J.; Tang, X.; Tang, X.; Ge, J.; Ding, B. Ultralight nanofibre-assembled cellular aerogels with superelasticity and multifunctionality. Nat. Commun. 2014, 5, 5802. [Google Scholar] [CrossRef] [Green Version]
  5. Si, Y.; Wang, X.; Dou, L.; Yu, J.; Ding, B. Ultralight and fire-resistant ceramic nanofibrous aerogels with temperature-invariant superelasticity. Sci. Adv. 2018, 4, eaas8925. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, H.; Zhang, X.; Wang, N.; Li, Y.; Feng, X.; Huag, Y.; Zhao, C.; Liu, Z.; Fang, M.; Ou, G.; et al. Ultralight, scalable, and high-temperature–resilient ceramic nanofiber sponges. Sci. Adv. 2017, 3, e1603170. [Google Scholar] [CrossRef] [Green Version]
  7. Li, Z.; Cui, Z.; Zhao, L.; Hussain, N.; Zhao, Y.; Yang, C.; Jiang, X.; Li, L.; Song, J.; Zhag, B.; et al. High-throughput production of kilogram-scale nanofibers by Kármán vortex solution blow spinning. Sci. Adv. 2022, 8, eabn3690. [Google Scholar] [CrossRef]
  8. George, G.; Luo, Z. A Review on Electrospun Luminescent Nanofibers: Photoluminescence Characteristics and Potential Applications. Curr. Nanosci. 2020, 16, 321–362. [Google Scholar] [CrossRef]
  9. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and Electrospun Nanofibers: Methods, Materials, and Applications. Chem. Rev. 2019, 119, 5298–5415. [Google Scholar] [CrossRef]
  10. Shamshirgar, A.S.; Rojas Hernández, R.E.; Tewari, G.C.; Fernández, J.F.; Ivanov, R.; Karppinen, M.; Hussainova, I. Functionally Graded Tunable Microwave Absorber with Graphene-Augmented Alumina Nanofibers. ACS Appl. Mater. Interfaces 2021, 13, 21613–21625. [Google Scholar] [CrossRef]
  11. Su, L.; Wang, H.; Niu, M.; Fan, X.; Ma, M.; Shi, Z.; Guo, S.-W. Ultralight, Recoverable, and High-Temperature-Resistant SiC Nanowire Aerogel. ACS Nano 2018, 12, 3103–3111. [Google Scholar] [CrossRef] [PubMed]
  12. Xue, Y.; Dai, P.; Zhou, M.; Wang, X.; Pakdel, A.; Zhang, C.; Weng, Q.; Takei, T.; Fu, X.; Popov, Z.; et al. Multifunctional superelastic foam-like boron nitride nanotubular cellular-network architectures. ACS Nano 2017, 11, 558–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Mahapatra, A.; Mishra, B.; Hota, G. Synthesis of ultra-fine α-Al2O3 fibers via electrospinning method. Ceram. Int. 2011, 37, 2329–2333. [Google Scholar] [CrossRef]
  14. Hou, T.; Li, X.; Lu, Y.; Yang, B. Highly porous fibers prepared by centrifugal spinning. Mater. Des. 2017, 114, 303–311. [Google Scholar] [CrossRef]
  15. Liu, H.-Y.; Chen, Y.; Liu, G.-S.; Pei, S.-G.; Liu, J.-Q.; Ji, H.; Wang, R.-D. Preparation of High-Quality Zirconia Fibers by Super-High Rotational Centrifugal Spinning of Inorganic Sol. Mater. Manuf. Process. 2013, 28, 133–138. [Google Scholar] [CrossRef]
  16. Tepekiran, B.N.; Calisir, M.D.; Polat, Y.; Akgul, Y.; Kilic, A. Centrifugally spun silica (SiO2) nanofibers for high-temperature air filtration. Aerosol Sci. Technol. 2019, 53, 921–932. [Google Scholar] [CrossRef]
  17. Sedaghat, A.; Taheri-Nassaj, E.; Naghizadeh, R. An alumina mat with a nano microstructure prepared by centrifugal spinning method. J. Non Cryst. Solids 2006, 352, 2818–2828. [Google Scholar] [CrossRef]
  18. Lu, Y.; Li, Y.; Zhang, S.; Xu, G.; Fu, K.; Lee, H.; Zhang, X. Parameter study and characterization for polyacrylonitrile nanofibers fabricated via centrifugal spinning process. Eur. Polym. J. 2013, 49, 3834–3845. [Google Scholar] [CrossRef]
  19. Wu, H.; Pan, W.; Lin, D.; Li, H. Electrospinning of ceramic nanofibers: Fabrication, assembly and applications. J. Adv. Ceram. 2012, 1, 2–23. [Google Scholar] [CrossRef] [Green Version]
  20. Xue, J.; Xie, J.; Liu, W.; Xia, Y. Electrospun Nanofibers: New Concepts, Materials, and Applications. Accounts Chem. Res. 2017, 50, 1976–1987. [Google Scholar] [CrossRef]
  21. Yan, J.; Zhao, Y.; Wang, X.; Xia, S.; Zhang, Y.; Han, Y.; Yu, J.; Ding, B. Polymer template synthesis of soft, light, and robust oxide ceramic films. Iscience 2019, 15, 185–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Song, X.; Liu, W.; Wang, J.; Xu, S.; Liu, B.; Cai, Q.; Ma, Y. Highly aligned continuous mullite nanofibers: Conjugate electrospinning fabrication, microstructure and mechanical properties. Mater. Lett. 2018, 212, 20–24. [Google Scholar] [CrossRef]
  23. Jia, C.; Li, L.; Song, J.; Li, Z.; Wu, H. Mass Production of Ultrafine Fibers by a Versatile Solution Blow Spinning Method. Accounts Mater. Res. 2021, 2, 432–446. [Google Scholar] [CrossRef]
  24. Daristotle, J.L.; Behrens, A.M.; Sandler, A.D.; Kofinas, P. A review of the fundamental principles and applications of solution blow spinning. ACS Appl. Mater. Interfaces 2016, 8, 34951–34963. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, H.; Liao, S.; Bai, X.; Liu, Z.; Fang, M.; Liu, T.; Wang, N.; Wu, H. Highly flexible indium tin oxide nanofiber transparent electrodes by blow spinning. ACS Appl. Mater. Interfaces 2016, 8, 32661–32666. [Google Scholar] [CrossRef]
  26. Rotta, M.; Zadorosny, L.; Carvalho, C.; Malmonge, J.; Malmonge, L. YBCO ceramic nanofibers obtained by the new technique of solution blow spinning. Ceram. Int. 2016, 42, 16230–16234. [Google Scholar] [CrossRef] [Green Version]
  27. Silva, V.D.; Ferreira, L.S.; Simões, T.A.; Medeiros, E.S.; Macedo, D.A. 1D hollow MFe2O4 (M=Cu, Co, Ni) fibers by Solution Blow Spinning for oxygen evolution reaction. J. Colloid Interface Sci. 2019, 540, 59–65. [Google Scholar] [CrossRef]
  28. Li, L.; Xu, C.; Chang, R.; Yang, C.; Jia, C.; Wang, L.; Song, J.; Li, Z.; Zhang, F.; Fang, B.; et al. Thermal-responsive, super-strong, ultrathin firewalls for quenching thermal runaway in high-energy battery modules. Energy Storage Mater. 2021, 40, 329–336. [Google Scholar] [CrossRef]
  29. Lu, B.; He, Y.; Duan, H.; Zhang, Y.; Li, X.; Zhu, C.; Xie, E. A new ultrahigh-speed method for the preparation of nanofibers containing living cells: A bridge towards industrial bioengineering applications. Nanoscale 2012, 4, 1003–1009. [Google Scholar] [CrossRef]
  30. Li, G.; Zhu, M.; Gong, W.; du, R.; Eychmüller, A.; Li, T.; Lv, W.; Zhang, X. Boron Nitride Aerogels with Super-Flexibility Ranging from Liquid Nitrogen Temperature to 1000 °C. Adv. Funct. Mater. 2019, 29, 1900188. [Google Scholar] [CrossRef]
  31. Ichikawa, H. Polymer-derived ceramic fibers. Annu. Rev. Mater. Sci. 2016, 46, 335–356. [Google Scholar] [CrossRef]
  32. Wang, F.; Dou, L.; Dai, J.; Li, Y.; Huang, L.; Si, Y.; Yu, J.; Ding, B. In situ Synthesis of Biomimetic Silica Nanofibrous Aerogels with Temperature-Invariant Superelasticity over One Million Compressions. Angew. Chem. 2020, 132, 8362–8369. [Google Scholar] [CrossRef]
  33. Wang, X.; Dou, L.; Li, Z.; Yang, L.; Yu, J.; Ding, B. Flexible hierarchical ZrO2 nanoparticle-embedded SiO2 nanofibrous membrane as a versatile tool for efficient removal of phosphate. ACS Appl. Mater. Interfaces 2016, 8, 34668–34676. [Google Scholar] [CrossRef]
  34. Mi, H.-Y.; Jing, X.; Huang, H.-X.; Turng, L.-S. Instantaneous self-assembly of three-dimensional silica fibers in electrospinning: Insights into fiber deposition behavior. Mater. Lett. 2017, 204, 45–48. [Google Scholar] [CrossRef]
  35. Mao, X.; Si, Y.; Chen, Y.; Yang, L.; Zhao, F.; Ding, B.; Yu, J. Silica nanofibrous membranes with robust flexibility and thermal stability for high-efficiency fine particulate filtration. RSC adv. 2012, 2, 12216–12223. [Google Scholar] [CrossRef]
  36. Huang, T.; Zhu, Y.; Zhu, J.; Yu, H.; Zhang, Q.; Zhu, M. Self-reinforcement of light, temperature-resistant silica nanofibrous aerogels with tunable mechanical properties. Adv. Fiber Mater. 2020, 2, 338–347. [Google Scholar] [CrossRef]
  37. Ren, L.; Simmons, T.J.; Lu, F.; Rahmi, O.; Kotha, S.P. Template free and large-scale fabrication of silica nanotubes with centrifugal jet spinning. Chem. Eng. J. 2014, 254, 39–45. [Google Scholar] [CrossRef]
  38. Ren, L.; Ozisik, R.; Kotha, S.P. Rapid and efficient fabrication of multilevel structured silica micro-/nanofibers by centrifugal jet spinning. J. Colloid Interface Sci. 2014, 425, 136–142. [Google Scholar] [CrossRef]
  39. Wang, Y.; Li, W.; Jiao, X.; Chen, D. Electrospinning preparation and adsorption properties of mesoporous alumina fibers. J. Mater. Chem. A 2013, 1, 10720–10726. [Google Scholar] [CrossRef]
  40. Wang, Y.; Li, W.; Xia, Y.; Jiao, X.; Chen, D. Electrospun flexible self-standing γ-alumina fibrous membranes and their potential as high-efficiency fine particulate filtration media. J. Mater. Chem. A 2014, 2, 15124–15131. [Google Scholar] [CrossRef]
  41. Song, X.; Zhang, K.; Song, Y.; Duan, Z.; Liu, Q.; Liu, Y. Morphology, microstructure and mechanical properties of electrospun alumina nanofibers prepared using different polymer templates: A comparative study. J. Alloys Compd. 2020, 829, 154502. [Google Scholar] [CrossRef]
  42. Li, L.; Kang, W.; Zhao, Y.; Li, Y.; Shi, J.; Cheng, B. Preparation of flexible ultra-fine Al2O3 fiber mats via the solution blowing method. Ceram. Int. 2015, 41, 409–415. [Google Scholar] [CrossRef]
  43. Akia, M.; Capitanachi, D.; Martinez, M.; Hernandez, C.; de Santiago, H.; Mao, Y.; Lozano, K. Development and optimization of alumina fine fibers utilizing a centrifugal spinning process. Microporous Mesoporous Mater. 2018, 262, 175–181. [Google Scholar] [CrossRef]
  44. Liu, R.; Dong, X.; Xie, S.; Jia, T.; Xue, Y.; Liu, J.; Jing, W.; Guo, A. Ultralight, thermal insulating, and high-temperature-resistant mullite-based nanofibrous aerogels. Chem. Eng. J. 2019, 360, 464–472. [Google Scholar] [CrossRef]
  45. Zadeh, M.M.A.; Keyanpour-Rad, M.; Ebadzadeh, T. Synthesis of mullite nanofibres by electrospinning of solutions containing different proportions of polyvinyl butyral. Ceram. Int. 2013, 39, 9079–9084. [Google Scholar] [CrossRef]
  46. Song, X.; Ma, Y.; Wang, J.; Liu, B.; Yao, S.; Cai, Q.; Liu, W. Homogeneous and flexible mullite nanofibers fabricated by electrospinning through diphasic mullite sol–gel route. J. Mater. Sci. 2018, 53, 14871–14883. [Google Scholar] [CrossRef]
  47. Xian, L.; Zhang, Y.; Wu, Y.; Zhang, X.; Dong, X.; Liu, J.; Guo, A. Microstructural evolution of mullite nanofibrous aerogels with different ice crystal growth inhibitors. Ceram. Int. 2020, 46, 1869–1875. [Google Scholar] [CrossRef]
  48. Farias, R.M.D.C.; Menezes, R.R.; Oliveira, J.E.; de Medeiros, E.S. Production of submicrometric fibers of mullite by solution blow spinning (SBS). Mater. Lett. 2015, 149, 47–49. [Google Scholar] [CrossRef]
  49. Ruiz-Rosas, R.; Bedia, J.; Rosas, J.; Lallave, M.; Loscertales, I.; Rodríguez-Mirasol, J.; Cordero, T. Methanol decomposition on electrospun zirconia nanofibers. Catal. Today 2012, 187, 77–87. [Google Scholar] [CrossRef]
  50. Chen, Y.; Mao, X.; Shan, H.; Yang, J.; Wang, H.; Chen, S.; Tian, F.; Yu, J.; Ding, B. Free-standing zirconia nanofibrous membranes with robust flexibility for corrosive liquid filtration. RSC Adv. 2014, 4, 2756–2763. [Google Scholar] [CrossRef]
  51. Wang, H.; Lin, S.; Yang, S.; Yang, X.; Song, J.; Wang, D.; Wang, H.; Liu, Z.; Li, B.; Fang, M.; et al. High-temperature particulate matter filtration with resilient yttria-stabilized ZrO2 nanofiber sponge. Small 2018, 14, 1800258. [Google Scholar] [CrossRef] [PubMed]
  52. Costa, D.L.; Leite, R.S.; Neves, G.A.; Santana, L.N.D.L.; Medeiros, E.S.; Menezes, R.R. Synthesis of TiO2 and ZnO nano and submicrometric fibers by solution blow spinning. Mater. Lett. 2016, 183, 109–113. [Google Scholar] [CrossRef]
  53. Wang, F.; Song, L.; Zhang, H.; Meng, Y.; Luo, L.; Xi, Y.; Liu, L.; Han, N.; Yang, Z.; Tang, J.; et al. ZnO Nanofiber Thin-Film Transistors with Low-Operating Voltages. Adv. Electron. Mater. 2017, 4, 1700336. [Google Scholar] [CrossRef] [Green Version]
  54. Ning, Y.; Zhang, Z.; Teng, F.; Fang, X. Novel Transparent and Self-Powered UV Photodetector Based on Crossed ZnO Nanofiber Array Homojunction. Small 2018, 14, e1703754. [Google Scholar] [CrossRef]
  55. Ahmad, M.; Pan, C.; Luo, Z.; Zhu, J. A single ZnO nanofiber-based highly sensitive amperometric glucose biosensor. J. Phys. Chem. C 2010, 114, 9308–9313. [Google Scholar] [CrossRef]
  56. Wang, X.; Zhao, M.; Liu, F.; Jia, J.; Li, X.; Cao, L. C2H2 gas sensor based on Ni-doped ZnO electrospun nanofibers. Ceram. Int. 2013, 39, 2883–2887. [Google Scholar] [CrossRef]
  57. Zhang, M.; Wang, Y.; Zhang, Y.; Song, J.; Si, Y.; Yan, J.; Ma, C.; Liu, Y.T.; Yu, J.; Ding, B. Conductive and elastic TiO2 nanofibrous aerogels: A new concept toward self-supported electrocatalysts with superior activity and durability. Angew. Chem. 2020, 59, 23252–23260. [Google Scholar] [CrossRef]
  58. Zhang, R.; Wang, X.; Song, J.; Si, Y.; Zhuang, X.; Yu, J.; Ding, B. In situ synthesis of flexible hierarchical TiO2 nanofibrous membranes with enhanced photocatalytic activity. J. Mater. Chem. A 2015, 3, 22136–22144. [Google Scholar] [CrossRef]
  59. Wu, M.-C.; Sápi, A.; Avila, A.; Szabó, M.; Hiltunen, J.; Huuhtanen, M.; Toth, G.; Kukovecz, A.; Konya, Z.; Keiski, R.; et al. Enhanced photocatalytic activity of TiO2 nanofibers and their flexible composite films: Decomposition of organic dyes and efficient H2 generation from ethanol-water mixtures. Nano Res. 2011, 4, 360–369. [Google Scholar] [CrossRef]
  60. Santos, A.; Mota, M.; Leite, R.; Neves, G.A.; Medeiros, E.; Menezes, R. Solution blow spun titania nanofibers from solutions of high inorganic/organic precursor ratio. Ceram. Int. 2018, 44, 1681–1689. [Google Scholar] [CrossRef]
  61. Gonzalez-Abrego, M.; Hernandez-Granados, A.; Guerrero-Bermea, C.; de la Cruz, A.M.; Garcia-Gutierrez, D.; Sepulveda-Guzman, S.; Cruz-Silva, R. Mesoporous titania nanofibers by solution blow spinning. J. Sol-Gel Sci. Technol. 2017, 81, 468–474. [Google Scholar] [CrossRef]
  62. Qian, L.; Willneff, E.; Zhang, H. A novel route to polymeric sub-micron fibers and their use as templates for inorganic structures. Chem. Commun. 2009, 3946–3948. [Google Scholar] [CrossRef] [PubMed]
  63. Korhonen, J.T.; Hiekkataipale, P.; Malm, J.; Karppinen, M.; Ikkala, O.; Ras, R.H.A. Inorganic Hollow Nanotube Aerogels by Atomic Layer Deposition onto Native Nanocellulose Templates. ACS Nano 2011, 5, 1967–1974. [Google Scholar] [CrossRef] [PubMed]
  64. Liu, H.; Chen, Y.; Pei, S.; Liu, G.; Liu, J. Preparation of nanocrystalline titanium dioxide fibers using sol–gel method and centrifugal spinning. J. Sol-Gel Sci. Technol. 2013, 65, 443–451. [Google Scholar] [CrossRef]
  65. Bao, N.; Wei, Z.; Ma, Z.; Liu, F.; Yin, G. Si-doped mesoporous TiO2 continuous fibers: Preparation by centrifugal spinning and photocatalytic properties. J. Hazard. Mater. 2010, 174, 129–136. [Google Scholar] [CrossRef]
  66. Dai, Y.; Tian, J.; Fu, W. Shape manipulation of porous CeO2 nanofibers: Facile fabrication, growth mechanism and catalytic elimination of soot particulates. J. Mater. Sci. 2019, 54, 10141–10152. [Google Scholar] [CrossRef]
  67. Wu, H.; Lin, D.; Pan, W. Fabrication, assembly, and electrical characterization of CuO nanofibers. Appl. Phys. Lett. 2006, 89, 133125. [Google Scholar] [CrossRef]
  68. An, Z.; Ye, C.; Zhang, R.; Zhou, P. Flexible and recoverable SiC nanofiber aerogels for electromagnetic wave absorption. Ceram. Int. 2019, 45, 22793–22801. [Google Scholar] [CrossRef]
  69. Liu, Y.; Liu, Y.; Choi, W.C.; Chae, S.; Lee, J.; Kim, B.-S.; Park, M.; Kim, H.Y. Highly flexible, erosion resistant and nitrogen doped hollow SiC fibrous mats for high temperature thermal insulators. J. Mater. Chem. A 2017, 5, 2664–2672. [Google Scholar] [CrossRef]
  70. Li, Y.; Yan, G.; Zhao, Y.; Kang, W.; Li, L.; Zhuang, X.; Cheng, B.; Li, X. Emulsion-blow spun self-sustained crystalline β-silicon carbide (SiC) fiber mat and its conductivity property. Trans. Indian Ceram. Soc. 2017, 76, 159–164. [Google Scholar] [CrossRef]
  71. Wang, Y.; Han, C.; Zheng, D.; Lei, Y. Large-scale, flexible and high-temperature resistant ZrO2/SiC ultrafine fibers with a radial gradient composition. J. Mater. Chem. A 2014, 2, 9607–9612. [Google Scholar] [CrossRef]
  72. Wu, N.; Wan, L.Y.; Wang, Y.; Ko, F. Conversion of hydrophilic SiOC nanofibrous membrane to robust hydrophobic materials by introducing palladium. Appl. Surf. Sci. 2017, 425, 750–757. [Google Scholar] [CrossRef]
  73. Guo, A.; Roso, M.; Modesti, M.; Liu, J.; Colombo, P. Preceramic polymer-derived SiOC fibers by electrospinning. J. Appl. Polym. Sci. 2014, 131, 39836. [Google Scholar] [CrossRef]
  74. Dong, X.; Liu, R.; Dong, W.; Wang, Z.; Guo, A.; Liu, J.; Chen, C.; Jiang, Y. Fabrication and properties of lightweight SiOC fiber-based assembly aerogels with hierarchical pore structure. Ceram. Int. 2018, 44, 22760–22766. [Google Scholar] [CrossRef]
  75. Wang, P.; Cheng, L.; Zhang, L. Lightweight, flexible SiCN ceramic nanowires applied as effective microwave absorbers in high frequency. Chem. Eng. J. 2018, 338, 248–260. [Google Scholar] [CrossRef]
  76. Song, Y.; Li, B.; Yang, S.; Ding, G.; Zhang, C.; Xie, X. Ultralight boron nitride aerogels via template-assisted chemical vapor deposition. Sci. Rep. 2015, 5, 10337. [Google Scholar] [CrossRef] [Green Version]
  77. Su, L.; Li, M.; Wang, H.; Niu, M.; Lu, D.; Cai, Z. Resilient Si3N4 nanobelt aerogel as fire-resistant and electromagnetic wave-transparent thermal insulator. ACS Appl. Mater. Interfaces 2019, 11, 15795–15803. [Google Scholar] [CrossRef]
  78. Huo, W.; Zhang, X.; Xu, J.; Hu, Z.; Yan, S.; Gan, K.; Yang, J. In situ synthesis of three-dimensional nanofiber-knitted ceramic foams via reactive sintering silicon foams. J. Am. Ceram. Soc. 2019, 102, 2245–2250. [Google Scholar] [CrossRef]
  79. Zhang, X.; Huo, W.; Chen, Y.; Hu, Z.; Wang, Y.; Gan, K.; Yang, J. Novel micro-spherical Si3N4 nanowire sponges from carbon-doped silica sol foams via reverse templating method. J. Am. Ceram. Soc. 2019, 102, 962–969. [Google Scholar]
  80. Wu, H.; Sun, Y.; Lin, D.; Zhang, R.; Zhang, C.; Pan, W. GaN nanofibers based on electrospinning: Facile synthesis, controlled assembly, precise doping, and application as high performance UV photodetector. Adv. Mater. 2009, 21, 227–231. [Google Scholar] [CrossRef]
  81. Dou, L.; Zhang, X.; Cheng, X.; Ma, Z.; Wang, X.; Si, Y.; Yu, J.; Ding, B. Hierarchical cellular structured ceramic nanofibrous aerogels with temperature-invariant superelasticity for thermal insulation. ACS Appl. Mater. Interfaces 2019, 11, 29056–29064. [Google Scholar] [CrossRef] [PubMed]
  82. Su, L.; Wang, H.; Niu, M.; Dai, S.; Cai, Z.; Yang, B.; Huyan, H.; Pan, X. Anisotropic and hierarchical SiC@ SiO2 nanowire aerogel with exceptional stiffness and stability for thermal superinsulation. Sci. Adv. 2020, 6, eaay6689. [Google Scholar]
  83. Xu, X.; Fu, S.; Guo, J.; Li, H.; Huang, Y.; Duan, X. Elastic ceramic aerogels for thermal superinsulation under extreme conditions. Mater. Today 2021, 42, 162–177. [Google Scholar] [CrossRef]
  84. Hu, P.; Liu, L.; Zhao, M.; Wang, J.; Ma, X.; Wang, J. Design, Synthesis, and Use of High Temperature Resistant Aerogels Exceeding 800 °C. ES Mater. Manuf. 2021, 15, 14–33. [Google Scholar] [CrossRef]
  85. Wang, B.; Niu, Y.; Zheng, S.; Yin, Y.; Ding, M. A High Temperature Sensor Based on Sapphire Fiber Fabry-Perot Interferometer. IEEE Photon Technol. Lett. 2019, 32, 89–92. [Google Scholar] [CrossRef]
  86. Chen, Y.; Chen, Y.; Ding, D.; Ding, Y.; Choi, Y.; Zhang, L.; Yoo, S.; Chen, D.; Deglee, B.; Xu, H.; et al. A robust and active hybrid catalyst for facile oxygen reduction in solid oxide fuel cells. Energy Environ. Sci. 2017, 10, 964–971. [Google Scholar] [CrossRef] [Green Version]
  87. He, Z.; Li, C.; Chen, C.; Tong, Y.; Luo, T.; Zhan, Z. Membrane-assisted propane partial oxidation for solid oxide fuel cell applications. J. Power Sources 2018, 392, 200–205. [Google Scholar] [CrossRef]
  88. Gao, Z.; Mogni, L.V.; Miller, E.C.; Railsback, J.G.; Barnett, S.A. A perspective on low-temperature solid oxide fuel cells. Energy Environ. Sci. 2016, 9, 1602–1644. [Google Scholar] [CrossRef]
  89. Yang, R.; Hu, F.; An, L.; Armstrong, J.N.; Hu, Y.; Li, C.; Huang, Y.; Ren, S. A Hierarchical Mesoporous Insulation Ceramic. Nano Lett. 2019, 20, 1110–1116. [Google Scholar] [CrossRef]
  90. Cheng, H.; Xue, H.; Hong, C.; Zhang, X. Preparation, mechanical, thermal and ablative properties of lightweight needled carbon fibre felt/phenolic resin aerogel composite with a bird’s nest structure. Compos. Sci. Technol. 2017, 140, 63–72. [Google Scholar] [CrossRef]
  91. Jin, H.; Zhou, X.; Xu, T.; Dai, C.; Gu, Y.; Yun, S.; Hu, T.; Guan, G.; Chen, J. Ultralight and Hydrophobic Palygorskite-based Aerogels with Prominent Thermal Insulation and Flame Retardancy. ACS Appl. Mater. Interfaces 2020, 12, 11815–11824. [Google Scholar] [CrossRef] [PubMed]
  92. Nguyen, B.N.; Meador, M.A.B.; Scheiman, D.; McCorkle, L. Polyimide Aerogels Using Triisocyanate as Cross-linker. ACS Appl. Mater. Interfaces 2017, 9, 27313–27321. [Google Scholar] [CrossRef] [PubMed]
  93. Randall, J.P.; Meador, M.A.B.; Jana, S.C. Tailoring Mechanical Properties of Aerogels for Aerospace Applications. ACS Appl. Mater. Interfaces 2011, 3, 613–626. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic diagram of (a) centrifugal spinning (Reprinted with permission from Ref. [18]. 2013 Elsevier) and (b) the equipment (Reprinted with permission from Ref. [14]. 2016 Elsevier).
Figure 1. Schematic diagram of (a) centrifugal spinning (Reprinted with permission from Ref. [18]. 2013 Elsevier) and (b) the equipment (Reprinted with permission from Ref. [14]. 2016 Elsevier).
Chemistry 04 00098 g001
Figure 2. Fabrication process of nanofibers (Reprinted with permission from Ref. [21]).
Figure 2. Fabrication process of nanofibers (Reprinted with permission from Ref. [21]).
Chemistry 04 00098 g002
Figure 3. Schematic diagram of conjugate electrospinning (Reprinted with permission from Ref. [22]. 2017 Elsevier).
Figure 3. Schematic diagram of conjugate electrospinning (Reprinted with permission from Ref. [22]. 2017 Elsevier).
Chemistry 04 00098 g003
Figure 4. Fabrication process of solution blow spinning (Reprinted with permission from Ref. [6]).
Figure 4. Fabrication process of solution blow spinning (Reprinted with permission from Ref. [6]).
Chemistry 04 00098 g004
Figure 5. High-pressure gas spinning. (A) The simple and novel droplet-elongating spinning instrument. The spray gun was connected to an air compressor. The nozzle facing the temperature field (60 °C) was equipped with a 200 mesh filter. (B) A high-speed digital image of the droplet-elongating spinning process in action. (C) Image of PVP nanofiber membrane of area 35 × 55 cm2 after 15 min spinning 450 mL PVP solution, showing the ultrahigh productivity of the extrusion spray spinning. (D) Scanning electron micrograph showing the uniform nanofibers distributed randomly. (E) Schematic illustration of the droplet-elongating spinning process: (a) pear-like droplet; (b) and (c) the tadpolelike droplet ‘‘wagging its tail’’ and its body being elongated; (d) small diameter fibers (Reprinted with permission from Ref. [29]. 2012 Royal Society of Chemistry).
Figure 5. High-pressure gas spinning. (A) The simple and novel droplet-elongating spinning instrument. The spray gun was connected to an air compressor. The nozzle facing the temperature field (60 °C) was equipped with a 200 mesh filter. (B) A high-speed digital image of the droplet-elongating spinning process in action. (C) Image of PVP nanofiber membrane of area 35 × 55 cm2 after 15 min spinning 450 mL PVP solution, showing the ultrahigh productivity of the extrusion spray spinning. (D) Scanning electron micrograph showing the uniform nanofibers distributed randomly. (E) Schematic illustration of the droplet-elongating spinning process: (a) pear-like droplet; (b) and (c) the tadpolelike droplet ‘‘wagging its tail’’ and its body being elongated; (d) small diameter fibers (Reprinted with permission from Ref. [29]. 2012 Royal Society of Chemistry).
Chemistry 04 00098 g005
Figure 6. Fabrication process of SiO2 nanofiber aerogel (CNFAs) (A) Schematic illustration of the fabrication of CNFAs. (B) XPS spectrum of CNFAs for all elements; a.u., arbitrary unit. (C) A CNFA heated by a butane blowtorch without any damage. (D) An optical image of CNFAs with diverse shapes. (E) An optical image showing a 20 cm3 CNFA (r = 0.15 mg·cm−3) standing on the tip of a feather. (FH) The microscopic structure of CNFAs at different magnifications demonstrates the hierarchical nanofibrous cellular architecture. (I) STEM-EDS images of a single nanofiber with corresponding elemental mapping images of Si, O, Al, and B, respectively. (J) Schematic showing the three levels of a hierarchy of the relevant structures. (Reprinted with permission from Ref. [5]).
Figure 6. Fabrication process of SiO2 nanofiber aerogel (CNFAs) (A) Schematic illustration of the fabrication of CNFAs. (B) XPS spectrum of CNFAs for all elements; a.u., arbitrary unit. (C) A CNFA heated by a butane blowtorch without any damage. (D) An optical image of CNFAs with diverse shapes. (E) An optical image showing a 20 cm3 CNFA (r = 0.15 mg·cm−3) standing on the tip of a feather. (FH) The microscopic structure of CNFAs at different magnifications demonstrates the hierarchical nanofibrous cellular architecture. (I) STEM-EDS images of a single nanofiber with corresponding elemental mapping images of Si, O, Al, and B, respectively. (J) Schematic showing the three levels of a hierarchy of the relevant structures. (Reprinted with permission from Ref. [5]).
Chemistry 04 00098 g006
Figure 7. Fiber forming mechanism via high-pressure gas spinning. (AC) show the particles’ track after they had moved for 0.005, 0.075 and 0.12 s, respectively. (D) Proposed mechanism to explain the formation of nanofibers in the extrusion spray spinning process. (i) Scanning electron micrographs of the PVP nanofiber (or droplet) collected at different locations from 3 to 90 cm away from the spray nozzle. The inset is a transmission electron micrograph of the PVP nanofibers collected 180 cm away from the spray nozzle. (ii) Schematic illustrations showing the development from the original PVP droplets to PVP nanofibers. (Reprinted with permission from Ref. [29]. 2012 Royal Society of Chemistry).
Figure 7. Fiber forming mechanism via high-pressure gas spinning. (AC) show the particles’ track after they had moved for 0.005, 0.075 and 0.12 s, respectively. (D) Proposed mechanism to explain the formation of nanofibers in the extrusion spray spinning process. (i) Scanning electron micrographs of the PVP nanofiber (or droplet) collected at different locations from 3 to 90 cm away from the spray nozzle. The inset is a transmission electron micrograph of the PVP nanofibers collected 180 cm away from the spray nozzle. (ii) Schematic illustrations showing the development from the original PVP droplets to PVP nanofibers. (Reprinted with permission from Ref. [29]. 2012 Royal Society of Chemistry).
Chemistry 04 00098 g007
Figure 8. The fabrication process of the mullite-based nanofibrous aerogel (Reprinted with permission from Ref. [44]. 2018 Elsevier).
Figure 8. The fabrication process of the mullite-based nanofibrous aerogel (Reprinted with permission from Ref. [44]. 2018 Elsevier).
Chemistry 04 00098 g008
Figure 9. Fabrication procedure illustration of N-doped hollow SiC fibrous mats (Reprinted with permission from Ref. [69]. 2013 Royal Society of Chemistry).
Figure 9. Fabrication procedure illustration of N-doped hollow SiC fibrous mats (Reprinted with permission from Ref. [69]. 2013 Royal Society of Chemistry).
Chemistry 04 00098 g009
Figure 10. The morphology of the electrospun SiOC fibers after being calcined at 1000 °C (a) the low resolution and (b) the high resolution; (c) the diameter distribution of the electrospun SiOC fibers after calcined at 1000 °C (Reprinted with permission from Ref. [71]. 2018 Elsevier).
Figure 10. The morphology of the electrospun SiOC fibers after being calcined at 1000 °C (a) the low resolution and (b) the high resolution; (c) the diameter distribution of the electrospun SiOC fibers after calcined at 1000 °C (Reprinted with permission from Ref. [71]. 2018 Elsevier).
Chemistry 04 00098 g010
Figure 11. Schematic drawing of chemical composition on the surface of SiOC and SiOCxPd (Reprinted with permission from Ref. [72]. 2017 Elsevier).
Figure 11. Schematic drawing of chemical composition on the surface of SiOC and SiOCxPd (Reprinted with permission from Ref. [72]. 2017 Elsevier).
Chemistry 04 00098 g011
Figure 12. BN nanofibers: (a,b) SEM images, (ce) TEM images, (f) an AFM image of BN nanoribbon aerogel, (g) STEM and element mappings of single BN nanoribbon, and (h) nitrogen adsorption–desorption isotherms of BN nanoribbon aerogels. Inset in (a), a BN nanoribbon aerogel monolith was placed on a flower. Inset in (b), the contact angle of BN aerogel. Inset in (d), the corresponding electron diffraction pattern. Inset in (h), the corresponding pore size distribution curve. (Reprinted with permission from Ref. [30]. 2019 WILEY-VCH Verlag GmbH & Co,. KGaA.)
Figure 12. BN nanofibers: (a,b) SEM images, (ce) TEM images, (f) an AFM image of BN nanoribbon aerogel, (g) STEM and element mappings of single BN nanoribbon, and (h) nitrogen adsorption–desorption isotherms of BN nanoribbon aerogels. Inset in (a), a BN nanoribbon aerogel monolith was placed on a flower. Inset in (b), the contact angle of BN aerogel. Inset in (d), the corresponding electron diffraction pattern. Inset in (h), the corresponding pore size distribution curve. (Reprinted with permission from Ref. [30]. 2019 WILEY-VCH Verlag GmbH & Co,. KGaA.)
Chemistry 04 00098 g012
Figure 13. Thermal insulation performance of the SiC NWA. (a) Fresh petal protected by a piece of aerogel with a thickness of 10 mm from withering or carbonization under the heating of an alcohol lamp for 10 min. (b) Thermal conductivity of the SiC NWA in N2 at different temperatures (Reprinted with permission from Ref. [11]. 2018 American Chemical Society).
Figure 13. Thermal insulation performance of the SiC NWA. (a) Fresh petal protected by a piece of aerogel with a thickness of 10 mm from withering or carbonization under the heating of an alcohol lamp for 10 min. (b) Thermal conductivity of the SiC NWA in N2 at different temperatures (Reprinted with permission from Ref. [11]. 2018 American Chemical Society).
Chemistry 04 00098 g013
Figure 14. (a) Fresh flower protected by HNA3 with a thickness of 2 cm under the heating of a butane blowtorch for 10 min. (b) Illustration of the measurement setup using a butane blowtorch. (c) Infrared thermal image of the front side subjected to the butane blowtorch flame. (d) Infrared images of the back side during the 30 min heating process: (d1) 30 s, (d2) 10 min, (d3) 20 min, and (d4) 30 min. (e) The time-dependent temperature profile of the center point on the backside. (f) Photograph of the front side after 30 min fire resistance test. (g) SEM of the front side after 30 min fire resistance test (Reprinted with permission from Ref. [81]. 2019 American Chemical Society).
Figure 14. (a) Fresh flower protected by HNA3 with a thickness of 2 cm under the heating of a butane blowtorch for 10 min. (b) Illustration of the measurement setup using a butane blowtorch. (c) Infrared thermal image of the front side subjected to the butane blowtorch flame. (d) Infrared images of the back side during the 30 min heating process: (d1) 30 s, (d2) 10 min, (d3) 20 min, and (d4) 30 min. (e) The time-dependent temperature profile of the center point on the backside. (f) Photograph of the front side after 30 min fire resistance test. (g) SEM of the front side after 30 min fire resistance test (Reprinted with permission from Ref. [81]. 2019 American Chemical Society).
Chemistry 04 00098 g014
Figure 15. Thermal superinsulation performance of the AH-SSCSNWA. Thermal conducting behavior of the AH-SSCSNWA in the (A) axial and (B) radial directions, respectively, showing the anisotropic heat transfer behavior in different directions. (C) The thermal conductivities of the AH-SSCSNWA in axial and radial directions. (D) A schematic illustration showing the mechanism to achieve thermal superinsulation. (Reprinted with permission from Ref. [82]).
Figure 15. Thermal superinsulation performance of the AH-SSCSNWA. Thermal conducting behavior of the AH-SSCSNWA in the (A) axial and (B) radial directions, respectively, showing the anisotropic heat transfer behavior in different directions. (C) The thermal conductivities of the AH-SSCSNWA in axial and radial directions. (D) A schematic illustration showing the mechanism to achieve thermal superinsulation. (Reprinted with permission from Ref. [82]).
Chemistry 04 00098 g015
Figure 16. Illustration of POX membrane reactor (Reprinted with permission from Ref. [87]. 2018 Elsevier).
Figure 16. Illustration of POX membrane reactor (Reprinted with permission from Ref. [87]. 2018 Elsevier).
Chemistry 04 00098 g016
Figure 17. (a) Thermal conductivities of the series of PGA foams dependent on average pore size and porosity. (b) Thermal conductivity corresponding to different post annealing temperature and thermal conductivity before and after hydrophobicity treatment (Reprinted with permission from Ref. [89]. 2020 American Chemical Society).
Figure 17. (a) Thermal conductivities of the series of PGA foams dependent on average pore size and porosity. (b) Thermal conductivity corresponding to different post annealing temperature and thermal conductivity before and after hydrophobicity treatment (Reprinted with permission from Ref. [89]. 2020 American Chemical Society).
Chemistry 04 00098 g017
Figure 18. Screenshot of sample NCF-PR1/4 during the ablation test in an arc-jet wind tunnel recorded by camera from 1 to 33 s (Reprinted with permission from Ref. [90]. 2017 Elsevier).
Figure 18. Screenshot of sample NCF-PR1/4 during the ablation test in an arc-jet wind tunnel recorded by camera from 1 to 33 s (Reprinted with permission from Ref. [90]. 2017 Elsevier).
Chemistry 04 00098 g018
Figure 19. Current and proposed aerospace applications of aerogels, showing (a) Mars rover, (b) an inflatable decelerator concept for EDL applications, and (c) an EVA suit (Reprinted with permission from Ref. [93]. 2011 American Chemical Society).
Figure 19. Current and proposed aerospace applications of aerogels, showing (a) Mars rover, (b) an inflatable decelerator concept for EDL applications, and (c) an EVA suit (Reprinted with permission from Ref. [93]. 2011 American Chemical Society).
Chemistry 04 00098 g019
Table 1. Comparison of the main method for nanofiber preparation [1] (Adapted with permission from Ref. [1]. 2022 Donghua University).
Table 1. Comparison of the main method for nanofiber preparation [1] (Adapted with permission from Ref. [1]. 2022 Donghua University).
MethodsMaterial TypesAdvantagesDisadvantagesFiber DiameterIndustrialization
Prospect
Centrifugal spinningMeltHigh efficiencyRelatively poor flexibility>1 μmHigh
SolutionHigh efficiency; broad material choiceRemoval of polymerTens of nanometers to a few micronsRelatively high
SolHigh efficiency; high yield; polymer freeUnuniform diameterHundreds of nanometers to tens of micronsRelatively high
ElectrospinningSolutionSmall and uniform diameter; diverse compositions and morphologiesLow efficiency; high voltage; removal of polymerTens of nanometers to a few micronsRelatively low
SolHigh yield; polymer freeNonuniform diameter; high voltageTens of nanometers to a few micronsRelatively low
Solution blow spinningSolutionSimple and safe process; high efficiency; diverse compositions and structuresRemoval of polymersTens of nanometers to a few micronsRelatively high
Self-assemblySolutionRibbon structureLimited material choice; low efficiency; removal of organicWidth: hundreds of nanometers to a few microns. Thickness: a few nanometersLow
Polymer conversionSolutionNon-oxide componentLow efficiency; calcination in inert atmosphereTens of nanometers to a few micronsRelatively low
Table 2. Physical properties of some high-temperature nanofibers aerogel.
Table 2. Physical properties of some high-temperature nanofibers aerogel.
MaterialsDensity
(mg·cm−1)
Thermal Conductivity
(W·m−1·K−1)
SiO2 nanofibrous [84]0.2~170.0223
Al2O3, SiO2 nanofiber [44]34.64~48.890.03274~0.04317
Al2O3, ZrO2 nanofibrous [84]-0.0322 at 1300 °C
Al2O3-SiO2 micron fibers and nanoparticles [84]0.360.082 at 1200 °C
BN nanobelts [30]15.50.0346 ± 0.0015
Si3N4 nanobelts [77]1.8~9.60.029
SiC Nanowires [84]110.025
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.; Zhang, J.; Sun, Q.; Ye, X.; Ma, X.; Wang, J. Sol–Gel Routes toward Ceramic Nanofibers for High-Performance Thermal Management. Chemistry 2022, 4, 1475-1497. https://doi.org/10.3390/chemistry4040098

AMA Style

Zhang J, Zhang J, Sun Q, Ye X, Ma X, Wang J. Sol–Gel Routes toward Ceramic Nanofibers for High-Performance Thermal Management. Chemistry. 2022; 4(4):1475-1497. https://doi.org/10.3390/chemistry4040098

Chicago/Turabian Style

Zhang, Jing, Junxiong Zhang, Qilong Sun, Xinli Ye, Xiaomin Ma, and Jin Wang. 2022. "Sol–Gel Routes toward Ceramic Nanofibers for High-Performance Thermal Management" Chemistry 4, no. 4: 1475-1497. https://doi.org/10.3390/chemistry4040098

Article Metrics

Back to TopTop