Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

A review on electrospun magnetic nanomaterials: methods, properties and applications

Yifan Jia a, Congyi Yang a, Xueyang Chen a, Wenqing Xue a, Helena J. Hutchins-Crawford b, Qianqian Yu *a, Paul D. Topham *b and Linge Wang *a
aSouth China Advanced Institute for Soft Matter Science and Technology, School of Molecular Science and Engineering, Guangdong Provincial Key Laboratory of Functional and Intelligent Hybrid Materials and Devices, South China University of Technology, Guangzhou 510640, China. E-mail: yuqianqian@scut.edu.cn; lingewang@scut.edu.cn
bChemical Engineering and Applied Chemistry, School of Infrastructure and Sustainable Engineering, College of Engineering and Physical Sciences, Aston University, Birmingham, B4 7ET, UK. E-mail: p.d.topham@aston.ac.uk

Received 31st March 2021 , Accepted 25th June 2021

First published on 28th June 2021


Abstract

Magnetic materials display attractive properties for a wide range of applications. More recently, interest has turned to significantly enhancing their behaviour for advanced technologies, by exploiting the remarkable advantages that nanoscale materials offer over their bulk counterparts. Electrospinning is a high-throughput method that can continuously produce nanoscale fibres, providing a versatile way to prepare novel magnetic nanomaterials. This article reviews 20 years of magnetic nanomaterials fabricated via electrospinning and introduces their two primary production methods: electrospinning polymer-based magnetic fibres directly from solution and electrospinning fibrous templates for post-treatment. Continual advances in electrospinning have enabled access to a variety of morphologies, which has led to magnetic materials having desirable flexibility, anisotropy and high specific surface area. Post-treatment methods, such as surface deposition, carbonization and calcination, further improve or even create unique magnetic properties in the materials. This renders them useful in broad ranging applications, including electromagnetic interference shielding (EMS), magnetic separation, tissue engineering scaffolding, hyperthermia treatment, drug delivery, nanogenerators and data storage. The processing methods of electrospun magnetic nanofibres, their properties and related applications are discussed throughout this review. Key areas for future research have been highlighted with the aim of stimulating advances in the development of electrospun magnetic nanomaterials for a wide range of applications.


1. Introduction

Materials that possess magnetic character often exhibit specific desirable properties. This has led to the inclusion of magnetic materials in an ever-growing range of applications, including hard magnetic materials within magneto-biology,1 magnetic medicine,2 magnetic separation3 and electrical machinery and soft magnetic materials in stator or rotator parts of generators and motors.4 In addition to the various types of magnetic materials, there is also a variety of magnetic functional materials with various niche functions and applications such as giant magnetic resistors, magneto-strictive materials, magnetic fluids and magnetic refrigeration materials. For these applications pure organic, pure inorganic and hybrid organic–inorganic composite materials have become of increasing interest.

Within the past few decades many classic bulk materials (such as magnetic materials) have been processed into shapes with one or more dimensions at the nanoscale, rendering the materials with desirable properties that nanotechnology delivers. Among these magnetic nanomaterials, materials with two-dimensional scale constraints such as nanofibres (NFs) achieve incredible advances due to their anisotropic nature. NFs exhibit great enhancement and control of many properties with the most notable being flexibility, large specific surface area, porosity and coercivity (Hc).

Electrospinning is a simple means of processing materials to form NFs, where polymer chains align themselves under an electrostatic force to form elongated, thin, filamentous nanostructures. During the process, a polymer solution (or melt) is stretched and deformed by the electrostatic force and a droplet forms at the tip of needle. The shape of the droplet is determined by gravity, viscosity, surface tension and electric field. In the process of electrospinning, the most common droplet shape is a cone, referred to as the Taylor cone.5,6 When charge repulsion exceeds surface tension the polymer is pulled from the end of the Taylor cone. The modes of flow are also determined by the aforementioned forces. The withdrawn flow initially experiences stable motion and is then forced into an unstable stage where the polymer solidifies in the air to form fibres, which are then received by the collector.

The final properties of the fibres are influenced by three key factors:

(i) solution properties (such as the viscosity, concentration, polymer molecular weight and dielectric properties of the solution);

(ii) processing parameters (such as applied voltage, needle-to-collector distance and feeding rate) and

(iii) environmental conditions (such as temperature, humidity and air flow around the system).

These factors demonstrate the diverse range of magnetic electrospun NFs that can be produced from a single chemical composition; as the structure of a single NF and the fibre assembly can be manipulated by other means. Indeed, advances in the control over electrospun NFs is envisioned to vastly benefit the magnetic material field. For example, electrospinning is the only known method used to prepare continuous ultra-long, thin NFs.7

To the best of our knowledge, there are no comprehensive reviews of electrospun magnetic NFs that evaluate their methods of production, properties and applications. Herein, we have reviewed articles from 20 years of research on electrospun magnetic fibres. The review is divided into three main parts (as shown in Fig. 1): (i) electrospinning organic–inorganic hybrid magnetic materials (summarised in Table 1); (ii) using electrospun fibres as templates for the creation of both hybrid and solely inorganic magnetic materials (summarised in Table 2); and (iii) applications of magnetic nanofibrous materials. As explained in this review, organic–inorganic composite magnetic nanomaterials can be prepared via a one-step method and the resulting nanofibrous matrix can provide magnetic nanoparticles (MNPs) with mechanical support, protection against oxidation and favourable dispersion (Section 2). Alternatively, for pure inorganic magnetic NFs (Section 3), electrospinning is a simple, available tool used in the fabrication of fibrous templates with different morphological structures enabling the user to manipulate the magnetic properties of the final product. The major difference between these two strategies is that organic matter is removed in the latter to create the final inorganic product. In the final section (Section 4), we explore the various applications of these advanced materials from electromagnetic interference (EMI) shielding and pollutant treatment, to biomedical devices in drug delivery and tissue engineering. In short, this review focuses on the processing methods of electrospun magnetic composite NFs and pure inorganic NFs from templates, their properties and related applications.


image file: d1tc01477c-f1.tif
Fig. 1 General methods to prepare electrospun magnetic fibres, where the area process highlighted in red is discussed in Section 2 (a) and the processes highlighted in blue are discussed in Section 3 (b and c). (a) Direct method for producing magnetic nanofibres; (b) templating procedure of MNFs from a magnetic pre-cursor solution and (c) templating procedure of MNFs where the magnetic component is deposited upon them.
Table 1 Organic–inorganic hybrid magnetic nanofibrous materials created directly from electrospinning solution(s) discussed in Section 2, detailing magnetic components and polymers used, electrospinning method followed, the magnetic properties of the resulting materials and their applications, where relevant
Magnetic ingredients Polymer Electrospinning approaches Magnetic properties Applications Ref.
Fe2O3 PVP Uniaxial – blend H c = 327 Oe, Mr/Ms = 0.244 44
PMMA, PU Uniaxial – blend Superparamagnetic, Ms = 6.172 emu g−1 37 and 90
PVA Uniaxial – deposition EWA 289
γ-Fe2O3 PLA Uniaxial – blend Paramagnetic, superparamagnetic, Ms = 0.049 emu g−1 Cell culture 88, 89, 254 and 255
Tissue engineering scaffolding
Oil adsorption/separation, cell culture
PVA Coaxial 290
Uniaxial – blend Paramagnetic, superparamagnetic Tissue engineering scaffolding 253
Fe3O4 CA Uniaxial – blend Superparamagnetic, Ms = 2.284 emu g−1 Water treatment 59
Cellulose–CS, PEO Uniaxial – blend Superparamagnetic, Ms = 18.61 emu g−1 (max) Fluorescence self-display and adsorption removal of mercury(II) 60
Cellulose Coaxial spinning – post treatment 291
Cellulose pulp Uniaxial – blend Ferromagnetic Nanofibrous scaffolds 292
CS Uniaxial – blend Superparamagnetic and ferromagnetic at different temperature Hyperthermia treatment of tumor cells 145
CS, PEO Uniaxial – blend Superparamagnetic, Ms = 8.47–7.12 emu g−1 Adsorption removal of heavy metal 293
Superparamagnetic, Ms = 11.21–16.94 emu g−1 (RT) Hypothermic tumor cell treatment 294
CMC, PVA Uniaxial – blend Ferromagnetic, Hc = 216–222 Oe 295
CS, PVA Uniaxial – blend Superparamagnetic, Ms = 0.67–3.19 emu g−1 Bone regeneration 296
CS[thin space (1/6-em)]:[thin space (1/6-em)]PVA = 3[thin space (1/6-em)]:[thin space (1/6-em)]7 Uniaxial – blend Superparamagnetic, Ms = 20.98 emu g−1 Chromium(VI) removal 22
CTMB Uniaxial – blend Enantioselective adsorption of racemic drug 297
DNA–CTMA Uniaxial – blend Superparamagnetic Water detoxification 298
Gelatin Uniaxial – blend Superparamagnetic, Ms = 3.05–12.87 emu g−1 20
HPMCP, CA Uniaxial – blend Superparamagnetic, Ms = 0.28–0.52 emu g−1 Drug release 18
MADO Uniaxial – blend Superparamagnetic, Ms = 83.9–74.6 emu g−1 Hypothermic chemotherapy 299
P(AN-co-AA) Uniaxial – blend Superparamagnetic, Ms = 27.02–30.51 emu g−1 300
PA6 Uniaxial – blend Superparamagnetic, Ms = 1.03 emu g−1 EMS 301 and 302
PAAm, PVA Uniaxial – blend Superparamagnetic, Ms = 16.6, 47.7, 48 emu g−1, Mr = 2.8 ± 0.5 emu g−1, Hc = 150 Oe 303
PAN Uniaxial – blend Superparamagnetic, Hc = 20.1–206.7 Oe, Ms = 4.67 emu g−1 Magnetic separation of the photosensitizers, electrets filter media 21 and 304–306
M s = 3.5–8.4 emu g−1 Cell separation, drug targeting 307
Ferromagnetic Microwave absorption 308
Superparamagnetic, Ms = 81.389 emu g−1 Phenol removal 38
M s = 8.6 emu g−1 (max) Separation of glycoproteins 309
Twisted blend Superparamagnetic, Ms = 10.10–28.77 emu g−1 33
Layer-by-layer M s = 5.56–20.76 emu g−1 36
PAN-co-AA Uniaxial – blend Oil adsorption/separation 310
Uniaxial – blend Superparamagnetic, Ms = 3.6 emu g−1 Adsorbents for removal of malachite green from water and wastewaters 311
PANI, PAN Layer-by-layer M s = 10.848–21.856 emu g−1 67
PANI, PVP Uniaxial – blend Superparamagnetic, Ms = 3.35–10.89 emu g−1 EMS 16
PANI/PMMA Coaxial M s = 4.52 emu g−1 66
PBT Uniaxial – blend Thin film microextraction, magnetic separation 312 and 313
PCL Uniaxial – blend Paramagnetic Magnetically-actuated, electromagnetic heating 314
Weak ferromagnetic or superparamagnetic, Hc ∼ 2.5 Oe, Mr ∼ 0.27 emu g−1, Ms = 1.0–11.2 emu g−1 Tissue engineering scaffold 256
M s = 27.7–103.9 emu g−1, Hc ∼ 70 Oe Organic pollutants degradation 315
Superparamagnetic Drug delivery vehicle 316
Ferrimagnetic, Ms = 88.5 emu g−1, Hc = 80 Oe Magnetic heating 58
Coaxial Superparamagnetic 317
Uniaxial – blend–UV cross-linking Superparamagnetic, Ms = 71.549 emu g−1 62
Coaxial Drug release 318
PCL[thin space (1/6-em)]:[thin space (1/6-em)]CS = 6[thin space (1/6-em)]:[thin space (1/6-em)]1 Uniaxial – blend Superparamagnetic, Ms = 0.74–3.52 emu g−1, Hc = 13.25–17.10 Oe Hyperthermia 319
PEK-C Uniaxial – blend–thermal treatment EWA 320
PEO Coaxial 42
PEO, PVA Janus Superparamagnetic 321
PEO/PLLA Uniaxial – blend Superparamagnetic Malachite green adsorption 322
PEO/PVP Uniaxial – blend 323
PET Uniaxial – blend M s = 0.58–2.79 emu g−1, Mr = 0.1–0.41 emu g−1, Hc = 79.94–103.9 Oe EMS 12
Ferromagnetic, near-superparamagnetic 71
Coaxial Superparamagnetic 43
PF–Na, PVA Uniaxial – blend Superparamagnetic, Ms = 9.7 emu g−1 324
PHB Uniaxial – blend M s = 2.4–4.9 emu g−1 Photocatalyst 325
PHB, PHVB Uniaxial – blend Superparamagnetic, Ms = 0.42–2.51 emu g−1 326
PHEMA, PLLA Uniaxial – blend Superparamagnetic 327
PLA Uniaxial – blend Paramagnetic Electromagnetic heating 31
PLA, PCL Uniaxial – blend Drug delivery 268
PLA, PEG Uniaxial – blend M s = 1.26–3.37 emu g−1 Smart clothing 328
PLGA Uniaxial – blend Superparamagnetic, Ms = 3.57–10.07 emu g−1 Tissue engineering scaffolds 85
PLLA Uniaxial – blend M s = 1.37–3.94 emu g−1, paramagnetic, superparamagnetic Tissue engineering scaffold 257 and 329
PMMA Uniaxial – blend Superparamagnetic, Ms = 5.22–23.19 emu g−1 55 and 330
Coaxial M r = 8.38 emu g−1, Ms = 4.23–35.77 emu g−1 63, 64, 74 and 331
Janus M s = 2.96–32.61 emu g−1, superparamagnetic 30, 65, 86 and 332–335
Janus – coaxial M s = 31.98 emu g−1 336
Uniaxial – blend – cospinning M s = 3.6–24.0 emu g−1 337
PMMA, PANI Uniaxial – blend – cospinning M s = 7.69 emu g−1 338
PMMA/PANI Janus Superparamagnetic, Ms = 23.52 emu g−1 28
PNIPAM Uniaxial – blend – cross-linking Superparamagnetic, Ms = 7.88 and 15.80 emu g−1 339
Polythiophene, CS Uniaxial – blend Solid-phase extraction of triazine herbicides 340
PS Uniaxial – blend Remote and efficient oil adsorption 57 and 341
Cancer therapy
PS, PVDF Two-nozzle blend Water oil separation 34
PVA Uniaxial – blend Superparamagnetic, Ms = 1.18–3.66 emu g−1, coercivity = 6.14–8.98 Oe, retentivity = 1.8–8.4 Oe, Ms = 2.77 emu g−1, Ms = 2.42 emu g−1 46, 51 and 342
Ferromagnetic, Ms = 26 emu g−1, Mr = 10 emu g−1, Hc = 20 Oe 61
Twisted blend Superparamagnetic, Ms = 7.11–21.28 emu g−1 56
PVA, guar gum Uniaxial – deposition Superparamagnetic, Ms = 0.1–5.8 emu g−1 94
PVA, PAA Uniaxial – blend Ferromagnetic, Ms = 1.72–6.77 emu g−1 72
PVA[thin space (1/6-em)]:[thin space (1/6-em)]PAA = 5[thin space (1/6-em)]:[thin space (1/6-em)]6 Uniaxial – blend Superparamagnetic, Ms = 10.9–39.9 emu g−1 Wastewater treatment 343
PVA, PCL Uniaxial – blend Tissue engineering scaffolds 258
PVA–PHB/PCL Coaxial M s = 1.9 ± 0.3 emu g−1 Photocatalyst 344
PVC Uniaxial – blend Microwave absorption 345
PVDF Uniaxial – blend Superparamagnetic, Ms = 1.93–12.5 emu g−1, Hc = 96–113 Oe, Ms = 46.5 emu g−1 Triboelectric nanogenerator 19, 282 and 346
Superparamagnetic Hyperthermia treatment and skin wound healing applications 347
PVP Coaxial Superparamagnetic, Ms = 1.76–13.59 emu g−1 26, 68, 70 and 348–350
Janus M s = 1.73–18.99 emu g−1 27, 35, 47, 87, 351 and 352
Superparamagnetic, Ms = 2.63–10.19 emu g−1 14, 69, 83, 84 and 353
Janus – uniaxial – blend M s = 7.4–14.84 emu g−1 354
Uniaxial – blend M s = 70.2 emu g−1, Mr = 8.7 emu g−1, Hc = 91 Oe EWA 355
Superparamagnetic, Ms = 36.6 emu g−1 356 and 357
PVP/PLLA Uniaxial – blend Superparamagnetic 73
Silk fibroin Uniaxial – blend & coating Superparamagnetic, Ms = 60 emu g−1 Tissue engineering scaffolds 8
PS-b-PI Coaxial Superparamagnetic (>13 kOe), Hc = 250 Oe (5 kOe) 91
Fe3O4, γ-Fe2O3 P(NIPAM-co-HMAAm) Uniaxial – blend & thermal crosslinking Induction of skin cancer apoptosis 259
Fe3O4, Fe2O3/NiO PAN Uniaxial – blend Data storage and transfer 1
Fe–FeO PI Uniaxial – blend M s = 30.6 emu g−1 (max), Hc = 188.2 Oe (max) 50
IONPs PLGA Uniaxial – blend Superparamagnetic, Ms = 18.84 emu g−1 Hyperthermia treatment and controlled drug release 358
PVP Uniaxial – blend Soft ferromagnetic, Ms = 0.53 emu g−1 266
PCL Uniaxial – blend Superparamagnetic, Ms = 6.8 emu g−1 Mesenchymal stem cell proliferation 359
SPIONs PDLLA Uniaxial – blend Superparamagnetic 360
CoFe2O4 PAN Uniaxial – blend & stabilization Superparamagnetic, Ms = 50 emu g−1, TB = 125 K 361
Nd0.05Bi0.95Fe0.95Co0.05O3 PVDF–TrFE Uniaxial – blend Ferromagnetic, Mr = 0.58 emu g−1, Hc = 1400 Oe 362
Magnetic bioglass PVA Uniaxial – blend Weak soft ferromagnetic, Hc ∼ 20 Oe, Mr ∼ 0.01 emu g−1 Bone scaffolds 260
FePt PCL Coaxial Superparamagnetic 101 and 363
Mg-ferrite PCL Uniaxial – blend Ferromagnetic, Ms = 0.024–3.19 emu g−1 Enhanced cell attachment, growth and proliferation 364
Magnetic zeolite PAN Uniaxial – blend M s = 15 emu g−1, coercivity = 97 Oe Determination of polycyclic aromatic hydrocarbons in water samples 365
Ni PS Uniaxial – blend M s = 0.08–1.52 emu g−1 98
SrFe12O19 PVA Uniaxial – blend M r/Ms = 0.72, Hc = 6.3 kOe Remove arsenic from water 17
Gd(DTPA) Eudragit S100, PEO Coaxial Drug delivery and MrI imaging 265
SrTiO3/NiFe2O4 (porous nanotubes & particle-in-tubes) PVP Uniaxial – blend & side-by-side uniaxial Ferromagnetic, Ms = 10–18 emu g−1 29
Fe-Doped In2O3/α-Fe2O3 PVP Coaxial M s = 0.48–22.37 emu g−1 366
CeO2-γ and CoFe2O4 PVP Uniaxial – blend H c = 320.40–541.35 Oe, Ms = 18.07–17.03 emu g−1, Mr = 3.72–4.04 emu g−1 Solar light driven photocatalyst 367
Mixture of magnetite (Fe3O4) and maghemite (γ-Fe2O3) P(NIPAM-co-HMAAm) Uniaxial – blend Inducing cancer apoptosis 1
SrRE0.6Fe11.4O19 (RE = La, Ce) PVP Sol–gel & uniaxial – blend H c = 4890.3–5321.9 Oe, Ms = 53.475–53.839 emu g−1, Mr = 28.517–28.765 emu g−1 368
BaFe12O19 PVP Uniaxial – blend & coaxial M s = 45.03–55.59 emu g−1, Mr = 1.2–27.12 emu g−1, Hc = 102.01–3761.57 Oe 369
CoFe2O4@Y2O3:5%Tb3 PVP Uniaxial & coaxial M s = 20.05–20.67 emu g−1 370
NaYF4:Eu3+ and Fe3O4 PVP Uniaxial – blend Superparamagnetic, Ms = 3.87–16.99 emu g−1 371
Fe PVP Uniaxial – blend 372
PANI Uniaxial – blend Magnetic hypothermia treatment 373
Co CA Uniaxial – blend Stem cells osteogenic differentiation 102
MnZnFe–Ni nanoparticles Estane Uniaxial – blend Superparamagnetic, Ms = 1.67–25 emu g−1 99
MGNPs Beta-lactoglobulin, PEO Uniaxial – blend M s = 0.1–4.16 emu g−1, Mr = 0.01–1.25 emu g−1, Hc = 89–114 Oe 78
Ni Polycarbonate-urethanes Uniaxial – blend 374
CoFe2O4, Fe3O4 Silk Uniaxial – blend Ferromagnetic 375
YFe garnet, YGdFe garnet CA Uniaxial – blend Bio-separation 376
Strontium hexaferrite PVA Uniaxial – blend Hard magnetic 76
TiO2/SiO2 PAN, PVP Uniaxial – blend EMS 377
MWCNTs PVA Uniaxial – blend EMS 378
PVP Uniaxial – blend EMS 379 and 380
FeCl3 PVA Uniaxial – blend 381
Ferritin PVA Uniaxial – blend Superparamagnetic Artificial muscles, MRI 100


Table 2 List of magnetic materials (both hybrid and solely inorganic) fabricated using a template approach discussed in Section 3, listing magnetic reagents, polymer(s) used, electrospinning and post-treatment procedures followed, morphology and properties of the final material obtained and applications where relevant
Magnetic ingredients Polymer Electrospinning approach and post-treatments Morphology (after post-treatment) Magnetic properties Applications Ref.
Fe PVP Uniaxial – carbonisation Regular M s = 85 emu g−1, Hc = 526 Oe EWA 232
Multi-nozzle – calcination – reduction Regular EWA 382
PAN Uniaxial – carbonisation Porous 3D cross-linked network Ferromagnetic, Ms = 32.95 emu g−1, Hc = 406 Oe EWA 226
Uniaxial – carbonisation–activation 3D non-woven network Ferromagnetic, Ms = 11.2–22.2 emu g−1 Catalyst for PMS activation 383
PVA Uniaxial – heating Regular H c = 0–275.0 Oe 120
Co PVP Uniaxial – stabilisation – carbonisation – mixed with paraffin and pressed into toroidal shaped specimens Regular Soft-magnetic, Ms = 73 emu g−1 EWA 227
Uniaxial – calcination – reduction Necklace-like Ferromagnetic, Ms = 28.37 emu g−1 (max), Hc = 674–1016 Oe 384
PVP & PAN Uniaxial – stabilisation – carbonisation Regular M s = 60 emu g−1 Catalysts 385
PAN Uniaxial – calcination Fibres with deposited particles Ferromagnetic, Ms = 10.1–23.9 emu g−1, Hc = 504.6–701.1 Oe Microwave absorption 386
Uniaxial – carbonisation Regular M s = 15 emu g−1 Catalyst for AR 387
TEOS/PVA Uniaxial – calcination – lyophilisation – reduction – coating Fibres with deposited particles M s = 27.1 emu g−1 (max) Absorbents 388
PVA Uniaxial – carbonisation Regular Ferromagnetic, Ms = 77.52 emu g−1, Hc = 261.3 Oe, Mr = 7.97 emu g−1, Hs = 8500 emu g−1 (RT), Ms = 78.45 emu g−1, Hc = 392.7 Oe, Mr = 9.26 emu g−1, Hs = 8500 emu g−1 (5 K) 389
Ni PVA Uniaxial – calcination Regular Ferromagnetic, Ms = 25.3 emu g−1, Hc = 382.52 Oe, Mr = 10.4 emu g−1, Hs = 3700 emu g−1 (5 K), Ms = 23.12 emu g−1, Hc = 67.66 Oe, Mr = 3.06 emu g−1, Hs = 800 emu g−1 (RT) 390
Uniaxial – calcination – deoxidation Regular M s = 51.9 emu g−1, Hc = 185 Oe, Mr = 16.9 emu g−1 127
PAN Uniaxial – calcination Porous Electromagnetic interference shielding 391
Uniaxial – carbonisation Regular Lithium-ion batteries 392
PAN & PDA Uniaxial – carbonisation – surface modification Short Biosensors 393
Fe3O4 PAN Uniaxial – stabilisation – carbonisation – mix with paraffin Regular M s = 3.32–13.53 emu g−1, Hc = 1.21–274 Oe, Mr = 0.05–1.8 emu g−1 EWA 196
Uniaxial – stabilisation – carbonisation Regular Ferromagnetic, Ms = 30 emu g−1, Mr = 2.69 emu g−1, Hc = 189 Oe EWA 197
Uniaxial – stabilisation – carbonisation – coating Core–shell M s = 4–7.5 emu g−1, Hc = 180–240 Oe, Mr = 1.49–2.62 emu g−1 EWA 150
Uniaxial – hydrothermal – deposition Nanofibres with deposited particles Adsorbents for water purification 112
Coaxial – carbonisation Regular M s = 10.18–39.65 emu g−1 161
Coaxial – carbonisation Core–shell M s = 6.1–81.2 emu g−1 Magnetic, electronic and bio-applications 15
PAN/BA-a Uniaxial – stabilisation – calcination Hierarchical porous H c = 71 Oe, Mr = 0.654 emu g−1 Absorbents for organic dyes in water 394
PAN/PMMA Uniaxial – calcination Porous nanobelts M s = 5.54–18.49 emu g−1 Absorbents for organic dyes 111
PAN–PEI Uniaxial – calcination Sintered particles to form a fibre Superparamagnetic, Ms = 78.79 emu g−1, Mr = 0.528–10.5 emu g−1 Absorbent 395
PVP Uniaxial – stabilisation – calcination – mix with bismaleimide Regular EWA 148
Uniaxial – calcination – reduction Necklace-like Ferromagnetic, Ms = 58.4 emu g−1, Hc = 186.7 Oe EWA 396
Uniaxial – calcination – reduction Smooth M s = 57.6 emu g−1, Hc = 188.4 Oe, Mr/Ms = 0.28 131
Uniaxial – oxygen plasma treatment Sintered particles to form a fibre Ferromagnetic, Ms = 72.4–35.51 emu g−1, Mr = 3.15–0.07 emu g−1, Hc = 24.59–1.35 Oe 41
PEO Uniaxial – calcination Short fibres with deposited nanoparticles EWA 397
PBZ Uniaxial – calcination Sintered particles to form a fibre M s = 5.95–9.22 emu g−1, superparamagnetic Water treatment 398
PVDF Uniaxial – deposition Nanofibres with clusters Remote controllable oil removal 399
PU Uniaxial – deposition Fibres with deposited particles Superparamagnetic, Ms = 33.12 emu g−1 Hyperthermia treatment 107
PEO & PVA Uniaxial – crosslinking – co-precipitation 3D cross-linked network with deposited particles Superparamagnetic, Ms = 9–18 emu g−1, TB = 70–75 K Hyperthermia treatment 109
CS Uniaxial – deposition 3D cross-linked network with deposited particles Superparamagnetic, Ms = 16.3–27.2 emu g−1 (300 K), ferromagnetic, Hc = 284–298 Oe, Mr = 4.9–7.9 emu g−1 (10 K) Hyperthermia treatment of tumour cells 145
PVA Uniaxial – carbonisation Regular Ferromagnetic, Ms = 50.27–62.8 emu g−1, Hc = 50.2–150.72 Oe 400
PVA Uniaxial – anneal – deposition 3D cross-link network 168
PVA/PAA Uniaxial – blend – thermal treatment – deposition Cross-linked mat with deposited particles Superparamagnetic, Ms = 32.5 emu g−1 Recyclable catalytic capacities 401
PAA Uniaxial – thermal treatment Fibres with deposited particles M s = 1.52–10.46 emu g−1 187
PANI Uniaxial – deposition Nanofibres with deposited particles Ferromagnetic, Ms = 1.9 emu g−1, Hc = 930 Oe, Mr = 25.83 × 10−3 emu g−1 108
α-Fe2O3 PVA Uniaxial – deposition – calcination Hollow Weak ferromagnetic, Ms = 24.4 emu g−1 Absorbents for dyes 162
Uniaxial – calcination Regular M s = 20 emu g−1, Hc = 40 Oe (α-Fe2O3) Catalyst for azo dyes degradation 245
Uniaxial – calcination Nanorod Superparamagnetic–ferromagnetic 188
PVP Uniaxial – calcination Nanotubes Ferromagnetic, Ms = 0–20 emu g−1, Hc = 150–760 Oe, permanent magnetic material Supercapacitor electrodes 217
Uniaxial – calcination Nanotubes Ferromagnetic, Hc = 256.71–628.18 Oe, Mr = 0.2872–0.4512 emu g−1 163
γ-Fe2O3 PLGA & PCL Uniaxial – deposition Nanofibres with deposited particles Superparamagnetic, Ms = 3.56 emu g−1 Stem cell differentiation 144
PVA Uniaxial – hydrothermal synthesis deposition – calcination Core–shell Ferromagnetic, Ms = 98.9 emu g−1, Hc = 175.5 Oe, Mr = 6.9 emu g−1 Sensors 137
Uniaxial – calcination Regular Ferromagnetic, Ms = 52.1 emu g−1, Hc = 460 Oe, Mr = 16.7 emu g−1, Hs = 6000 Oe (5 K); Ms = 45.2 emu g−1, Hc = 218 Oe, Mr = 11 emu g−1, Hs = 2000 Oe (RT) Semiconductor 402
PVP Uniaxial – calcination Core–shell fibres, fibre-in-tube, tube-in-tube Ferromagnetic, Ms = 55.2 emu g−1 (fibre-in-tube), Ms = 56.3 emu g−1 (tube-in-tube), Mr = 5.5–15.9 emu g−1, Hc = 78–206 Oe, Mr/Ms = 0.1–0.28 403
Fe2O3 PVP Uniaxial – calcination Regular Superparamagnetic–antiferromagnetic Catalyst 177
Uniaxial – calcination – post treatment Porous hollow fibres with nanoflakes coating M s = 0.6 emu g−1 Photocatalyst 237
Uniaxial electrospinning – calcination Sintered particles to form a fibre Superparamagnetic, Ms = 8.42 emu g−1, Mr = 1.2 emu g−1, Hc = 160 Oe Photocatalyst 236
FexOy PVP Uniaxial – calcination Porous nanosheets, nanotubes Ferromagnetic, Ms = 2.84–18.91 emu g−1, Hc = 106.27–152.87 Oe 105
α-Fe2O3, Fe3O4 PAA Uniaxial electrospinning – deposition Nanofibres with deposited particles Superparamagnetic, Ms = 2.8–4.0 emu g−1 143
α-Fe2O3, Co3O4 PVP Uniaxial – calcination Hollow M s = 0.75–1.50 emu g−1 103
Co3O4 PAN Uniaxial – stabilisation – carbonisation Regular Soft ferromagnetic EWA 121
NiO PVP Uniaxial – calcination Sintered particles to form a fibre M s = 55.3 emu g−1, Hc = 194.0 Oe 404
PEtOx Uniaxial – calcination Sintered particles to form a fibre Alcohol sensor 405
Uniaxial – calcination Regular Ferromagnetic, Hc = 107.3 Oe (max), Mr = 0.472 emu g−1 (max) 203
CuO/NiO PVA Uniaxial – calcination Regular Paramagnetic, M = 0.337–0.480 emu g−1, Hc = 10 kOe 406
ZnO PVA Uniaxial – anneal Regular Ferromagnetic, Ms = 0.039 emu g−1 (max) 192
ZrO2 PVP Coaxial electrospinning – anneal – deposition Hollow Catalysts 166
Uniaxial – anneal Nanofibres with smooth surface M s = 0.45–0.57 emu g−1 Photocatalysts for the degradation of organic pollutes 241
SnO2 PVP Uniaxial – anneal Nanotubes M s = 0.012–0.017 emu g−1, Hc = 79–95 Oe (300 K), Ms = 0.11–0.19 emu g−1, Hc = 163–169 Oe (5 K) 173
Fe3C PAN Uniaxial – carbonisation Short Soft ferromagnetic, Ms = 18.0 emu g−1, Hc = 108.3 Oe, Mr = 0.68 emu g−1 EWA 228
Fe3O4, α-Fe2O3, Fe2N PVP Uniaxial – calcination Sintered particles to form a hollow fibre Fe3O4: Ms = 82.99 emu g−1, Mr = 39.27 emu g−1, Hc = 400.45 Oe α-Fe2O3: Ms = 4.34 emu g−1, Fe2N: Ms 2.07 emu g−1 High performance anodes for LIBs 134
FeC3/FeN3 PVP Uniaxial – stabilisation – nitridation Branch-like (400–700 °C) Ferromagnetic, Ms = 122 emu g−1 Fe, Hc = 112 Oe Fe, Mr = 10.4 emu g−1 Fe 407
Fe@FeO PAN Uniaxial – stabilisation – carbonisation Short Ferromagnetic, Ms = 6.8–24.1 emu g−1, Hc = 45–209 Oe 408
Pd doped Co PVA Uniaxial – calcination Regular Photocatalyst 409
Fe-Doped NiO PVA Uniaxial electrospinning – calcination Regular Ferromagnetic Diluted magnetic semiconductor 223
Fe doped ZnO PVA Uniaxial – calcination Regular T c > 300 K 410
Co doped ZnO PVP Uniaxial – calcination Regular Ferromagnetic, Ms = 0.05 emu g−1, Hc = 53.2 Oe, magnetic loss factor = 0.170–0.223 EWA 411
Fe doped SnO2/TiO2 PVP Uniaxial – calcination Beaded fibres Ferromagnetic, Ms = 0.02–0.37 emu g−1 Photocatalyst 412
Mn doped SnO2 PVP Uniaxial – calcination Hollow Ferromagnetic, paramagnetic, Hc = 15 kOe 221
Cu doped SnO2 PVP Coaxial – calcination Hollow RT ferromagnetism 413
GO doped CoFe2O4 PVP Uniaxial electrospinning – calcination Regular Ferrimagnetic, Ms = 79.24–82.7 emu g−1, Hc = 909–1514 Oe, Mr = 31–39 emu g−1 224
La-Doped TiO2/CoFe2O4 PVP (Sol–gel) – two-spinneret Regular M s = 8.888 emu g−1 Photocatalytic 414
Co-Doped SrTiO3 PVP Uniaxial – calcination – anneal Regular Paramagnetic–weak ferromagnetic 129
Gd doped bismuth ferrite PVP Uniaxial – annealing Regular M s = 2.4–4.12 emu g−1, Hc = 450 Oe 415
Ca doped BiFeO3 PVP Uniaxial – calcination Sintered particles to form a fibre Ferromagnetic Photocatalyst 240
Bi2O3 doped Ni0.5Zn0.5Fe2O4 PVP Uniaxial – calcination Regular M s = 2.6–59.1 emu g−1, Hc = 32.6–112.9 Oe 416
FexCoy PAN/PBZ Uniaxial – activation – carbonisation Regular M s = 18.76 emu g−1 Catalyst for PMS activation 417
PVA Uniaxial – graphinisation Fibres encapsulated in graphite shell Ferromagnetic, Ms = 71.14 emu g−1, Hc = 220 Oe (300 K), Hc = 648 Oe (5 K) 225
Fe–Ni PVP Uniaxial – calcination – deoxidation Regular Ferromagnetic, Ms = 72.54–195.06 emu g−1, Hc = 1.72–43.89 Oe 128
Uniaxial – calcination – deoxidation Nanoribbons Soft magnetic, Ms = 145.7 emu g−1 (max), Hc = 132 Oe 126
Fe3Si PVP Uniaxial – calcination Fibres with deposited particles Ferromagnetic, Ms = 0–8.4 emu g−1, Hc = 50–90 Oe Tunable EM and microwave absorption 418
FePt PVP Uniaxial – calcination – reduction process Necklace-like sintered particles to form a fibre Hard magnetic, Ms = 54.63–59.38 emu g−1, Hc = 4.68–10.27 Oe, Mr = 24.44–34.23 emu g−1 123
CoNi PVA Uniaxial – calcination Regular Ferromagnetic, Ms = 47.45 emu g−1, Hc = 65.6 Oe, Mr = 5.94 emu g−1, Hs = 4000 Oe 419
CoPt PVA Uniaxial – carbonisation Bead Ferromagnetic, Ms = 77.3 emu g−1 (Co, max), Hc = 270.9 Oe (Co–Pd, max), Mr = 7.98 emu g−1 (Co, min) 420
Sm2Co17 PVP Uniaxial – blend – calcination Regular M s = 55.5–106 emu g−1, Mr = 35.6–52.5 emu g−1, Hc = 5210–12[thin space (1/6-em)]676 Oe 206
CoxFeyAl PVP & PVA Uniaxial – anneal Regular Ferromagnetic 421
SmCoFe PVP Uniaxial – calcination – REDOX post – treatment Sintered particles to form a fibre M s = 80–120 emu g−1, Hc = 5–7.5 kOe, Mr = 55–69 emu g−1, Mr/Ms = 0.54–0.69 422
Co–MnO PVA Uniaxial – calcination Regular Ferromagnetic, Ms = 49.95 emu g−1, Hc = 245 Oe, Mr = 7.35 emu g−1, Hs = 3000 emu g−1 423
Zn1−xCoxO PVP Uniaxial – calcination – anneal Regular RT ferromagnetic, Ms = 0.877 emu g−1, Hc = 610 Oe (max) 424
PVA Uniaxial – sinter Regular H c = 50–75 Oe 218
Zn–Mn–O TPEE Uniaxial – calcination Microsphere composed of micro/nanofibres Ferromagnetic, Ms = 0.20225–0.78425 emu g−1, Hc = 83.68–223.78 Oe, Mr = 0.010125–0.015766 emu g−1 Photocatalyst 425
Fe2O3, Fe3O4, CuFe2O4, Cu2Fe2O4 PAN Coaxial – calcination Porous Ferromagnetic, Ms = 2.058 emu g−1, Mr = 0.28148 emu g−1, Hc = 167.56 Oe EWA 426
Magnetically susceptible conjugation complex Uniaxial – deposition – surface modification Regular Biosensor 427
CoFe2O4 PVP Uniaxial – anneal Nanotubes M s = 18 emu g−1 Photocatalysts 164
Uniaxial – calcination – (in situ) oxidative polymerisation method Hollow core–double shell nanostructure Photocatalysts 158
Uniaxial – calcination Wrinkle nanofibres with cluster M s = 33.232 emu g−1, Hc = 893.71 Oe, Mr = 10.876 emu g−1 Photocatalyst 200
Uniaxial – calcination Nanorod with flake surface Ferromagnetism, Ms = 35.17–61.24 emu g−1 Photocatalyst 202
Uniaxial – calcination – redox post – treatment Regular Efficient catalysts for the p-nitrophenol hydrogenation 428
Uniaxial – calcination Sintered particles to form a fibre 429
Uniaxial – calcination Sintered particles to form a fibre M s = 53.2–71.7 emu g−1, Hc = 925.3–1161.7 Oe 106
Uniaxial – anneal Janus M s = 41.34 emu g−1 75
Uniaxial – calcination Sintered particles to form a fibre Soft magnetic, superparamagnetic, Ms = 28.61–67.24 emu g−1, Hc = 758.62–2221.51 Oe, Mr = 8.28–35.11 emu g−1 172
Uniaxial – calcination Sintered particles to form a fibre Ferromagnetic, Ms = 42.8 emu g−1, Hc (bulk) = 750–1000 Oe, Mr/Ms = 0.27–0.5 430
(Dual-channel) – calcination Sintered particles to form two phases that are distributed semi-cylindrically H c = 250 Oe 431
Uniaxial – calcination Nanotubes H c = 300 Oe (360 K), Hc = 10[thin space (1/6-em)]400 Oe (5 K) 432
Uniaxial – ultrasonic – coaxial Regular M s = 3.65–45.80 emu g−1, Hc = 735–785 Oe, Mr = 1.4–16.49 emu g−1 433
Uniaxial – calcination Nanoribbons ferromagnetic, Ms = 64.6–80.3 emu g−1, Hc = 1223–1802 Oe 119
Uniaxial – calcination Nanoribbons (400–700 °C), nanofibres (900 °C) M s = 76.2–85.2 emu g−1 (2–300 K), Hc = 895 Oe (max), Mr/Ms = 0.75–0.89 114
Uniaxial – calcination Hollow Ferromagnetic, Ms = 9.0–34.7 emu g−1, Hc = 858–1207 Oe, Mr = 2.7–13.1 emu g−1, Mr/Ms = 0.30–0.38 (300 K), Ms = 9.6–36.1 emu g−1, Hc = 11[thin space (1/6-em)]953–14[thin space (1/6-em)]110 Oe, Mr = 8.3–32.5 emu g−1, Mr/Ms = 0.86–0.89 (2 K) 160
Uniaxial – calcination Wire-in-tube structure H c = 11[thin space (1/6-em)]043 Oe (10 K), Hc = 707 Oe (300 K) 434
Coaxial – anneal Core–shell M s = 16.1 emu g−1 156
Uniaxial – coprecipitate – calcination Nanofibres with deposited particles Photocatalyst 242
Uniaxial – anneal Porous nanoribbons M s = 56 emu g−1, Hc = 757 Oe, Mr = 18 emu g−1 (300 K), Ms = 72 emu g−1, Hc = 14[thin space (1/6-em)]507 Oe, Mr = 56 emu g−1 (2 K) 115
PAN Uniaxial – carbonisation Regular M s = 30 emu g−1 Catalyst for PMs activation 243
Uniaxial – stabilisation – carbonisation Regular M s = 50–63 emu g−1, Hc = 0–667 Oe, Mr = 0–17 emu g−1 94
PANI Uniaxial – calcination – polyaniline assisted (self-assembly) process – deposition Hollow fibres with deposited particles M s = 35 emu g−1, Hc = 1260 Oe Catalysts 244
PVAc (Sol–gel) uniaxial – calcination Regular M r = 16.1–32.5 emu g−1, Hc = 611.9–786.5 Oe 194
PVDF Uniaxial – blend – calcined – casted Regular M s = 10.7 emu cm−3 435
CoFe2O4/barium carbonate PVP Uniaxial – calcination – electrical assembly Janus Ferromagnetic, Ms = 60 emu g−1, Hc = 800 Oe, Mr = 18 emu g−1 Magnetoelectric sensors 436
CoFe2O4/Ag PVP Uniaxial – calcination Sintered particles to form a hollow fibre Catalyst for the degradation of organic pollutants 235
CoFe2O4–Pb(Zr0.52Ti0.48)O3 PS Uniaxial – anneal Regular Ferromagnetic, Hc = 386–730 Oe, Mr = 3.3–11.3 emu g−1 437
PVP & PMMA Coaxial – anneal Core–shell Ferromagnetic, Hc = 700 Oe, Mr = 3.40 emu g−1 140
CoFe2O4/CoFe2 PVP Uniaxial – calcination – partially reduction Regular Ferromagnetic, hard magnetic, Ms = 66.8–220.2 emu g−1, Hc = 0.62–1.37 Oe, Mr = 27.6–106.4 emu g−1 124
CoFe2O4/SrFe12O19 PVP Uniaxial – calcination Hollow M s = 52.9–62.8 emu g−1, Hc = 1089–4046 Oe, Mr = 18.22–31.21 emu g−1 438
CoFe2O4, NiFe2O4 PVP Coaxial – calcination Core–shell, fibre-in-tube, tube-in-tube M s = 63.83–76.16 emu g−1, Hc = 12.79–13.36 kOe 439
ZnFe2O4/CoFe2O4 PVP Uniaxial – calcination Short M s = 53.95–69.62 emu g−1, Hc = 69–110 Oe, Mr = 3.8–11 emu g−1, Mr/Ms = 0.07043–0.15799 440
Mullite fibres with embedded Ni NPs AN & PEO Uniaxial – thermal reduction – heat treatment Regular Ferromagnetic, Ms = 0.085–4.177 emu g−1, Hc = 21.1–85.8 Oe, Mr = 0.005–1.136 emu g−1 132
Fe–Ni/NiFe2O4 PVP Uniaxial – calcination – partially reduction Regular Soft magnetic, Ms = 49.5–109.3 emu g−1, Hc = 219–460 Oe 136
NiFe2O4 & MWCNTs PAN Uniaxial – stabilisation – carbonisation Regular Ferromagnetic, Ms = 1.5 emu g−1, Hc = 47 Oe EM shielding 142
Spinel-NiMn2O4 PVP Uniaxial – anneal Short fibres Paramagnetic–antiferromagnetic (200–300 K), Ms = 34 emu g−1, Hc = 96.7 Oe, Mr = 13 emu g−1 (5 K) Electrode material and energy storage 175
NiFe2O4 PVP Uniaxial – calcination – atomic layer deposition Core–shell Photocatalyst 441
Uniaxial – calcination Multi-particle-chain-like M s = 2.7 × 108 emu cm−3, Hc = 166 Oe, Tc = 858 K 442
Uniaxial – calcination Sintered particles to form a fibre Soft ferromagnetic, Ms = 1.3–40 emu g−1, Hc = 16–170 Oe, Mr = 0.008–7.8 emu g−1 183
Uniaxial – calcination Regular H c = 30 Oe (300 K), Hc = 576 Oe (5 K) 443
Uniaxial – anneal Nanotube Ferromagnetic, Ms = 33.3 emu g−1, Hc = 225 Oe 444
PVA Uniaxial – anneal Regular Ferromagnetic, Hc = 60 Oe 445
PVA & TEOS Uniaxial – (dip-coating) – calcination Hierarchical porous cross-linked structure Soft magnetic, Ms = 14.44 emu g−1 110
NiCo2O4 PAN Uniaxial – stabilisation – carbonisation Regular EWA 229
Ni0.8Co0.2Fe2O4, Ni PVP, PAN Uniaxial – carbonisation – mix with paraffin Regular M s = 11.9–53.4 emu g−1, Hc = 114.3–409 Oe EWA 446
Co0.5Ni0.5Fe2O4 PVP Uniaxial – calcination Necklace-like M s = 37.5–66.2 emu g−1, Hc = 78.3 Oe (max) EWA 122
Co1−xNixFe2O4 (x = 0.0, 0.2, 0.4, 0.6, 0.8, 1.0) PVP Uniaxial – stabilisation – calcination Regular M s = 13.4–56.4 emu g−1, Hc = 210–1255 Oe 215
MnFe2O4 PVP Uniaxial – calcination Short fibres Catalysts (magnetic separation) 447
Uniaxial – calcination Nanorods M s = 43.5–46.3 emu g−1 174
PVAc Uniaxial – calcination Regular M = 46.8–61.14 emu g−1 (10 kOe), Hc = 607.7–642.6 Oe, Mr = 13.4–17.8 emu g−1 (600–800 K) Magnetic recording device, magneto-optical recording and electronic devices 186
PAN Uniaxial – carbonisation Regular M s = 56.8 emu g−1 (max), Hc = 998–1264 Oe, Mr = 15–31.9 emu g−1, Mr/Ms = 0.41–0.67 185
Co0.5Mn0.5Fe2O4 PVP Uniaxial – calcination – ANI polymerisation Hollow Photocatalyst 448
CuFe2O4 PVP Uniaxial – calcination Hollow Ferromagnetic, Ms = 18.99–25.04 emu g−1, Hc = 345–1451 Oe, Mr = 5.32–13.02 emu g−1, Mr/Ms = 0.28–0.52 159
Uniaxial – deposition Nano fibres with deposited particles Visible light photocatalyst (magnetic separation) 199
Uniaxial – calcination Sintered particles to form a fibre or lamellar post sintering Ferromagnetic, soft magnetic, Ms = 7.73–23.98 emu g−1, Hc = 299–625 Oe, Mr = 6.08–9.91 emu g−1, Mr/Ms = 0.34–0.41 182
Uniaxial – calcination Sintered particles to form short or hollow fibres Ferromagnetic CR adsorption and CO catalytic oxidation of the samples 449
CuCo2O4 PAN Uniaxial – calcination Regular Ferromagnetism – antiferromagnetism, Ms = 11.43–60.24 (×10−3) emu g−1, Hc = 107.07–734.67 Oe, Mr = 4.02–25.78 (×10−3) emu g−1, Mr/Ms = 0.289–0.428 190
Ni1−xCuxFe2O4 PVP Uniaxial – calcination Regular Soft ferromagnetic, Ms = 15.1–37.71 emu g−1, Hc = 50.9–144.66 Oe 210
ZnFe2O4 PVP Uniaxial – anneal Porous nanotubes Photocatalyst 450
Uniaxial – calcination Sintered particles to form a fibre Ferromagnetic, Ms = 12.4 emu g−1, Hc = 48.79 Oe EWA 451
Uniaxial – calcination Sintered particles to form a fibre Superparamagnetic – paramagnetic (calcined at 500–700 °C), Ms = 1.53–2.55 emu g−1 193
ZnFe2O4/ZnO PVP Uniaxial – calcination Porous nanotubes Photocatalyst 452
ZnFe2O4/Fe3O4 PVP Uniaxial – calcination Regular Ferromagnetic, Ms = 22.9–34.2 emu g−1, Hc = 208.7–425.1 Oe, Mr = 5.8–10.1 emu g−1 Photocatalysts for wastewater disposal or drug release 453
ZnFe2O4/γ-Fe2O3 PVA Uniaxial – calcination Nanoribbons Ferromagnetic, Ms = 45 emu g−1 (max) 454
Cu1−xZnxFe2O4 PVP Uniaxial – calcination Regular Ferromagnetic–paramagnetic, Ms = 58.4 emu g−1 (max), Hc = 35.2–723.5 Oe, Mr/Ms = 0.11–0.47 emu g−1 216
Co0.6Zn0.4Fe2O4 PVP Uniaxial – calcination Regular Ferromagnetic, Ms = 92.3 emu g−1 (max), Hc = 338.2 Oe (max) EWA 455
Ni0.5−xCuxZn0.5Fe2O4 PVP Uniaxial – calcination Regular H c = 121.6 Oe (298 K), Hc = 295.9 Oe (77 K) 211
Ni0.5Zn0.5Fe2O4 PVP Uniaxial – anneal – calcination Nanowires in nanotubes M s = 61 emu g−1, Hc = 43 Oe, Mr = 3 emu g−1 456
Uniaxial – calcination Regular Soft magnetic, Ms = 78.6 emu g−1, Hc = 57.4 Oe EWA 457
NiZn ferrite PVP Uniaxial – calcination Sintered particles to form a fibre Ferromagnetic DNA separation for clinical diagnoses and biomolecular recognition 458
MgFe2O4 PVP Uniaxial – calcination Sintered particles to form a fibre M s = 17–31.1 emu g−1, Hc = 35.8–98.9 Oe 184
Uniaxial – anneal – sinter Nanotubes Ferromagnetic 135
Mg1−xZnxFe2O4 PVP Uniaxial – calcination Regular M s = 20.25 emu g−1, Mr = 5.1 emu g−1, Hc = 90 Oe, soft ferromagnetic 459
CuAl0.95Co0.05O2 PVP Uniaxial – anneal Regular Ferromagnetic, Ms = 0.012 emu g−1, Hc = 65.26 Oe, Mr = 0.0015 emu g−1 104
Fe@TiSi P123, PVP Coaxial – calcination Core–shell M s = 0.0002–1.2 emu g−1 Catalyst for wastewater treatment 249
α-Fe2O3@SiO2 TEOS Uniaxial – calcination – deposition Core–shell M s = 14–20 emu g−1, Hc = 390–400 Oe 151
SiO2–CoFe2O4 PVP & TEOS Uniaxial – anneal Hollow short fibres Ferromagnetic, Ms = 56.4–80 emu g−1, Hc = 1477 Oe 201
CaFe2O4 PVP Uniaxial – calcination Necklace-like Superparamagnetic Photocatalyst 238
CaFe2O4/MgFe2O4 PVP Uniaxial – calcination Sintered particles to form fibres Photocatalyst 460
TiO2/CoFe2O4 PVP (Sol–gel) – vertical two – spinneret – calcination Regular H c = 585.09 Oe, Mr = 3.2408 emu g−1, Ms = 9.5869 emu g−1 Photocatalyst 461
Ti0.9V0.1O2 PVP Uniaxial – calcination Regular Ferromagnetic 191
Cr0.046Zn0.954O PVP Uniaxial – calcination Regular Weak ferrimagnetic, Hc = 73–224 Oe 462
SrFe12O19 PVP Coaxial – stabilisation – sinter Hollow M s = 30.7–57.8 emu g−1, Hc = 95.3–4187.1 Oe EWA 167
Uniaxial – calcination – deposition – calcination Core–sheath Ferromagnetic, Ms = 7.968 emu g−1, Hc = 5149.7 Oe, Mr = 4.235 emu g−1 Photocatalyst 3
Uniaxial – calcination Necklace-like M s = 64 emu g−1, Hc = 6533.3 Oe High density magnetic recording and microwave devices 138
Uniaxial – anneal Nanoribbons M s = 50–67.9 emu g−1, Hc = 7150–7310 Oe, Mr = 27.5–37.3 emu g−1 116
Uniaxial – calcination Necklace-like M s = 59.9–60.8 emu g−1, Hc = 4538.2–6565.9 Oe 176
SrFe12O19/FeCo PVP Coaxial – calcination – reduction Core–shell M s = 60.9–68.8 emu g−1, Hc = 1249–3190 Oe 153
SrAlxFe12−xO19 (x = 0–3.0) PVP Uniaxial – calcination Necklace-like M s = 13–62 emu g−1, Hc = 5668–7737 Oe (298 K), Ms = 16–85 emu g−1, Hc = 6860 Oe (77 K) 209
SrTi1−xCoxO3 PVP Uniaxial – calcination – anneal Regular Ferromagnetic, Ms = 0.74 emu g−1 (max) 130
SrTi1−xFexO3 (SrTi0.9Fe0.1O3) PVP Uniaxial – calcination Regular Paramagnetic–ferromagnetic, M = 0.46–0.82 emu g−1, Hc = 165–217 Oe, Mr = 0.01–0.15 emu g−1, Tc > 300 K Batteries 463
SrTiO3/SrFe12O19 PVP Uniaxial – calcination Regular Hard magnetic, Ms = 55.1 emu g−1, Hc = 6523 Oe (max), Mr = 5.94 emu g−1 181
SrFe12O19, Ni0.5Zn0.5Fe2O4 PVP Uniaxial – calcination Sintered particles to form a fibre H c = 3037.1–4762.7 Oe (77–297 K), Mr = 53.5–39.1 emu g−1 (77–297 K) 464
Uniaxial – calcination Sintered particles to form a fibre M s = 56.1–64.9 emu g−1, Hc = 1484.7 Oe, Mr = 31.5 emu g−1 465
xSrSiO3/(100 − x)SrFe12O19 PVP Uniaxial – calcination Necklace-like M s = 45.6–58.0 emu g−1, Hc = 6283.8 Oe (max) 466
Yttrium iron garnet PVP Uniaxial – presintering – calcination Sintered particles to form a fibre M s = 3.1–21.5 emu g−1, Hc = 21.5–140 Oe 170
BaFe12O19 PVP Coaxial – calcination Hollow M s = 46.15–51.56 emu g−1, Hc = 5226 Oe, Mr = 22.54–27.59 emu g−1 467
Uniaxial – calcination Sintered particles to form a hollow fibre M s = 17.8 emu g−1, Hc = 4106.9 Oe 468
Uniaxial – calcination Sintered particles to form a fibre M s = 71.5 emu g−1, Hc = 5943 Oe 469
BaFe12−xAlxO19 PVP Uniaxial – calcination Sintered particles to form a fibre Hard magnetic, Ms = 29.70–63.92 emu g−1, Hc = 3614.0–9288.4 Oe, Mr = 17.73–33.44 emu g−1, Mr/Ms = 0.52–0.60 212
BaTi0.90Mn0.10O3 PVP Uniaxial – anneal Sintered particles to form a fibre Paramagnetic 222
Ni0.4Co0.2Zn0.4Fe2O4/BaTiO3 PVP Uniaxial – stabilisation – calcination Necklace-like M s = 34.5–64.4 emu g−1, Hc = 93.2–132.8 Oe EWA 470
Fe2O3, BaTiO3 PAN Uniaxial – stabilisation – carbonisation Regular Superparamagnetic, Ms = 12–12.5 emu g−1, Hc = 87–95 Oe Electromagnetic interference shields 471
Ba0.7Sr0.3TiO3–Ni0.8Zn0.2Fe2O4 PVP Uniaxial – anneal Regular Ferromagnetic, Ms = 2.479 emu g−1, Hc = 44.873 Oe, Mr = 0.189 emu g−1 472
Cobalt ferrite/barium calcium titanate PVP Uniaxial – magnetic field – sinter Regular Soft magnetic 473
Nickel ferrite/barium titanate PVP Coaxial – anneal Core–shell 48
LaFeO3 PVP Uniaxial – calcination Short fibres Ferromagnetic, Hc = 28[thin space (1/6-em)]078 Oe, Mr = 0.23 emu g−1 171
LaMnO3+δ PVA Uniaxial – calcination Null T c = 255 K, TB = 180 K 474
La1−xSrxMnO3 PVP Uniaxial – calcination Regular Ferromagnetic, Tc ≈ 365 K 475
La0.5Sr0.5TiO3 PVP Uniaxial – calcination Regular Ferromagnetic, Ms = 0.022–0.067 emu g−1 476
La0.7Sr0.3MnO3 PVP Uniaxial – calcination Regular M s = 1.23–40.52 emu g−1, Hc = 62 Oe (max) 179
Uniaxial – calcination Hollow Superparamagnetic–ferromagnetic, Ms = 1.8–50 emu g−1 (max), 180
Ba1−xLaxFe12O19 PVP Uniaxial – calcination Regular M s = 77.188 emu g−1, Hc = 3559.0 Oe (min) EWA 208
Sr0.8La0.2Zn0.2Fe11.8O19 PVP Uniaxial – calcination Necklace-like M s = 43.2–72.2 emu g−1, Hc = 164.3–5478.7 Oe 477
Sr1−xLaxFe12−xCoxO19 (x = 0.12) PVP Uniaxial – sinter Regular M s = 70.76 emu g−1, Hc = 6.26 kOe, Mr = 36.35 emu g−1 478
Ce0.96Fe0.04O2 PVP Uniaxial – calcination Null Ferromagnetic, Tc > 390 K, Ms = 0.0025–0.0923 emu g−1, Hc = 10 kOe 220
CuFe2O4@CeO2 PVP Uniaxial – calcination – precipitation – calcination Core–shell M s = 20.51–28.32 emu g−1, Mr = 7.24–12.85 emu g−1 Dye removal 152
SnO2/Ce PVP Uniaxial – calcination Porous hollow Ferromagnetic, Ms = 19 × 10−5 emu g−1 479
La0.33Pr0.34Ca0.33MnO3 PVP Uniaxial – calcination Regular T c ≈ 150 K, TB ≈ 50 K Magnetoresistance 219
Hard Sm2Co17 core and soft Fe11Co5 shell PVP Uniaxial – calcination – soft magnet plating Core–shell Permanent magnet 154
α-NaYF4:Yb/Er/Gd PEI Uniaxial – anneal Hollow Drug delivery 165
Na(Y/Gd)F4:Yb3+,Er3+ PVP Uniaxial – anneal Porous Drug delivery 480
GdOF:Er3+ PVP Uniaxial – calcination – fluorination – oxidation Regular Paramagnetic, mass magnetic susceptibility = 1.4252 × 10−4 emu g−1 Oe−1 (±20 Oe) 481
GdF3:Eu3+ PVP Uniaxial – calcination – fluorination Regular Paramagnetic, magnetisation, Ms = 2.11–2.46 emu g−1 482
NaGdF4:Dy3+ PVP Uniaxial – calcination – fluorination Nanobelts Superparamagnetic, Ms = 34.17–56.13 emu g−1, Hc = −50 to 50 kOe, mass magnetic susceptibility = 2.2239–3.7438 × 10−5 emu g−1 Oe−1 125
NaGdF4:0.5%Dy3+,Eu3+ PVP Uniaxial – calcination – fluorination Regular Magnetisation = 1.69–2.00 emu g−1 483
Gd2O2S:Dy3+,Eu3+ PVP Uniaxial – sinter – sulfurising Regular M s = 2.56–2.68 emu g−1 484
SrDyxFe12−xO19 PVP Uniaxial – calcination Regular Hard magnetic, Ms = 53–64 emu g−1, Hc = 6835–7155 Oe 155
Ni0.8Gd0.2Fe2O4 PVP Uniaxial – calcination Nanoribbon M s = 29.45–36.71 emu g−1, Hc = 8.7–33.73 Oe, Mr = 0.52–1.93 emu g−1 213
Co0.5Cu0.2Ni0.2Zn0.1Fe2O4, Co0.94Gd0.06Fe2O4.03 PVP Uniaxial – calcination Sintered particles to form a fibre Soft magnetic, Ms = 33.32–67.37 emu g−1 (Co0.5Cu0.2Ni0.2Zn0.1Fe2O4), Ms = 12.44–71.98 emu g−1 (Co0.94Gd0.06Fe2O4.03) 214
BiFeO3 PVP Uniaxial – calcination Regular Weak ferromagnetic Photocatalyst 45
Uniaxial – stabilisation – anneal Regular Ferromagnetic, good magnetic response, Ms = 4.4 emu g−1, Hc = 170 Oe Photocatalyst 239
Uniaxial – calcination – hydrothermal method Nanofibres with nanosheet deposited Ferromagnetic Photocatalyst 485
Uniaxial – anneal Short Weak-ferromagnetic 139
Uniaxial – calcination Necklace-like Magnetic moment = 0.4–3.0 emu g−1 486
Uniaxial – anneal Nanotubes Ferromagnetic, Tc = 300 K 113
Uniaxial – anneal Regular Weak ferromagnetic, Ms = 4 emu g−1 (max), Hc = 200 Oe 487
Uniaxial – calcination Regular Ferromagnetic, Ms = 4 emu g−1 207
Uniaxial – calcination – grind – Uniaxial Regular Ferromagnetic, Ms = 0.18–0.36 emu g−1 488
Nylon-6 Uniaxial – calcination Regular Ferromagnetic 489
Ni0.5Zn0.5Fe2O4/Pb(Zr0.52Ti0.48)O3 PVP Uniaxial – calcination Regular Ferroelectric, soft magnetic 178
Bi5Ti3FeO15 PVP Uniaxial – calcination Compact pellet to form a fibre Weak ferromagnetic, Ms = 6.7 × 10−4 emu g−1, Hc = 255 Oe, Mr = 0.92 × 10−4 emu g−1, Mr/Ms = 0.14 490
Bi2Sr2CaCu2O8+x PVP Uniaxial – anneal – calcination Regular T c = 78.7 K 491
Bi0.95Y0.05FeO3 PVP Sol–gel uniaxial – calcination Sintered particles to form a fibre M = 1.143–1.996 emu g−1 (MH = 2 T) 492
Bi0.9La0.1Fe0.95Mn0.05O3 PVP Uniaxial – calcination Regular M s = 404.7 memu g−1, Hc = 687.72 Oe, Mr = 48.115 memu g−1 493
Li7La3Zr2O12 PVP Uniaxial – calcination Regular 494
Nd0.1Bi0.9FeO3 PVP Sol–gel – uniaxial – carbonisation Hollow Antiferromagnetic, Ms = 1.82 emu g−1, Hc = 300 Oe 495
PbZr0.52Ti0.48O3–NiFe2O4 PVP Coaxial – anneal – (self-assembly) Core–shell Magnetisation = 6.5–9.6 emu g−1 157
Fe3O4@SiO2 PVP Uniaxial – blend – heated – hydrothermal treatment Fibres with deposited particles Superparamagnetic Photocatalyst 496
Montmorillonite PVDF Uniaxial – blend Agglomerates EWA 497


Finally, as aforementioned, we summarise the complete literature of magnetic NFs in two comprehensive tables, where Table 1 covers systems in the category described in Section 2 [i.e. organic–inorganic hybrid magnetic nanofibrous materials created directly from electrospinning solution(s)] and Table 2 lists those systems created from electrospun fibrous templates, as described in Section 3. Both tables describe the methodological approaches used, properties obtained and potential applications, where relevant. We anticipate that these tables will serve as a useful repository for researchers, in addition to those new to the field, looking to study such advanced materials.

2. Electrospinning magnetic materials: reagents, methods and morphologies

Electrospun polymer–magnetic material composite fibres have been widely reported across the literature. This section discusses the various electrospinning methods and parameters used to manipulate the creation of magnetic NFs, before reviewing the variety of magnetic materials that have been used. Table 1 serves as a comprehensive summary of this section, listing the various systems that have been explored, the magnetic reagent(s), polymer(s), electrospinning approach, major magnetic property and applications (where relevant) in each case.

2.1 Electrospinning methods and fibre structures

Electrospinning is a primary method used to prepare continuous nanofibrous materials with advantages such as simple device manufacture, material compatibility and controllable fibre morphology. Fibres produced by electrospinning have high surface area-to-volume ratios, tuneable surface morphology and controlled alignment. The specific morphology of the fibre can control the distribution of nanoparticles (NPs).8 In this section, we discuss the various electrospinning parameters that can be used to create an array of NFs, specifically focusing on magnetic NFs.
2.1.1 Materials. NFs can be produced by electrospinning from a solution or melt of a variety of materials. Adequate entanglement is the key for the material to be suitable for electrospinning. Most polymeric materials produce enough entanglement to perform both solution and melt electrospinning. The extent of entanglement is related to the polymer molecular weight, the concentration and the solvent system. Sufficiently low molecular weight, low concentration or poor solubility typically result in beaded structures (beaded fibres or discrete microspheres) rather than long defect-free fibres. Wang et al.9 studied the effect of solvent on the preparation of fibres and beads and found that decreasing the compatibility between the solvent and the polymer decreased the intermolecular entanglement, thus producing beads.

Nanoparticles can also be electrospun by adding nanoscale components into the polymer carrier. However, in order to form uniform nanoparticles some materials require pre-treatment. Common examples of pre-treatment include sol–gel treatment of the polymer solution and dispersive pre-treatment of composite nanoparticles, such as ultrasonic dispersion10–13 or coating with oleic acid (or grafting with a coupling agent) to prevent undesirable aggregation.14–17

2.1.2 Electrospinning rigs. Fig. 2 shows the wide range of electrospinning rigs that have been used to create various nanofibrous fabrics. The most commonly used electrospinning rig is relatively simple and is comprised of a power supply, a liquid supply device and a collector (Fig. 2(a)). The power supply provides sufficient electric field force to create liquid filaments. Adjustment of the liquid supply parameters such as; flow rate of solution, needle structure and mode of liquid supply, allow for continuous and uniform fibres to be formed. In addition to common single and multiple needle set-ups, there are also needleless devices such as silk thread, bulged roller and liquid pool. In these cases, the collector is usually a flat plate and the morphology of the fibre mat can be regulated by changing the collection method.11
image file: d1tc01477c-f2.tif
Fig. 2 Variety of different electrospinning apparatus including: (a) three basic parts, (b) core–shell needle, (c) multi-channel needle, (d) Janus fibre needle, (e) fibre blend, (f) layer-by-layer fibre, (g) drum collector, (h) twisted fibre collector, (i) multi-needle electrospinning, (j) drum electrospinning and (k) magnetically assisted electrospinning.

2.1.2.1 Uniaxial electrospinning. Uniaxial electrospinning is the most simplistic device design (Fig. 2(a)). The liquid supply uses a solitary nozzle to prepare a single component of solid micro/nanofibre.18 Fibres have been prepared, via uniaxial electrospinning, with different morphologies including; nanobelt (Fig. 3(c)), bead-on-string (Fig. 3(d)), and connected fibre mesh structures.19–22 The various morphologies provide different characteristics, such as hydrophobicity and mechanical properties. Superhydrophobicity is caused by the huge specific surface area of the fibrous mesh membrane. The large surface area of the membrane can greatly reduce the contact area between the fibrous membrane surface and the liquid. Additionally, the micro- and nanoscale voids on the surface of the specimen can easily trap air, and when water droplets come into contact with the material the air holds the droplets up in accordance with the typical Cassie–Baxster contact model.23 Additionally, it has been shown that electrospinning can be used to construct a composite structure of electrospun nanofibres decorated with microspheres to provide the superhydrophobic character. On this basis it becomes easier to design and control the micro–nano structure of the surface, allowing the perfect realisation of superhydrophobic structures.24,25
image file: d1tc01477c-f3.tif
Fig. 3 SEM images of electrospun fibres and particles: (a) NFs with encapsulated MNPs;8 (b) NFs with dip-coated MNPs;8 (c) nanobelts;55 (d) beaded fibres;19 (e) hollow fibres;26 (f) Janus NFs;14 (g) cross section of a bi-layered composite nanofibrous film;36 (h) the fibre of left layer containing the terbium complex Tb(TTA)3(TPPO)2 (where TTA is thenoyltrifluoroacetone and TPPO is triphenylphosphine oxide);36 (i) the fibre of right layer containing PANI·Fe3O4·PAN;36 reproduced from ref. 36 with permission from the PCCP Owner Societies; (j) random fibre mat;45 (k) oriented fibre mat;45 reproduced from ref. 45 with permission from the PCCP Owner Societies (l) yarn twist fibres56 (Published by The Royal Society of Chemistry); and (m–q) electrosprayed and electrospun fibres of styrene–(ethylene-co-butylene)–styrene from neat tetrahydrofuran (THF) solutions at varying polymer concentration (8 wt%, 10 wt%, 12 wt%, 14 wt% and 18 wt% for m–q, respectively).9

2.1.2.2 Coaxial electrospinning. Coaxial electrospinning (Fig. 2(b and c)) can be used to produce continuous, single-channel, or multi-channel core–shell and hollow fibres. The fibres are prepared by using two or more coaxial nozzles of differing diameter, which are loaded with the shell and core materials, respectively. The high speed of the jets prevents the disparate materials from mixing, resulting in a distinct boundary between the core and shell materials. Coaxial electrospinning is a useful method for separating the material of the inner and outer layers whilst protecting the load material (typically contained within the core). Yu et al.26 loaded air into the inner syringe to prepare hollow NFs (Fig. 3(e)) to be applied in targeted drug delivery applications.
2.1.2.3 Janus structure electrospinning. Janus nanostructures consist of two segregated materials with distinct physical and chemical properties to create a single nanostructure with two ‘faces’. Commonly, the two materials exhibit antagonistic properties, such as being hydrophilic/hydrophobic (polar/nonpolar), which forms an important area of research in materials science. To create Janus NFs, two different fibres are produced from separate nozzles that have opposite charges. The two different fibres attract one another to form the final Janus structure (Fig. 3(f)). The Janus structure stops the two parts of the material interfering with one another to prevent the loss of dual performance. This technology has been utilised to produce materials with superior properties in catalysis, sensing, biomedicine and display technology.27–30
2.1.2.4 Collecting methods and collectors. In addition to nozzle design, and the whip and curing stages of the electrospinning process, the overall morphology of the fibre can be controlled by changing the method of collection. In addition to a stationary plate, common collection methods include (but are not limited to) additional magnetic field,13,31,32 mechanical traction,33 spinning34 and layer-by-layer blending35,36 (Fig. 2). For example, Yang et al.35 prepared a novel sandwich-structured pellicle via layer-by-layer blending where the electric, magnetic and luminous layers could be effectively isolated from each other.

The arrangement of fibres can also be controlled by using different collectors. Again, compared to the more traditional flat plate, other collectors include the drum37–41 (Fig. 2(g)), parallel roller11,42–44 (Fig. 2(h)) and slit collector. These collectors allow NF membranes to be obtained with intricate patterns (Fig. 3(j–l)) for more innovative applications.38,45 Among these collectors, the drum is the most common for preparing oriented fibre fabrics and membranes.46


2.1.2.5 Combinations. The aforementioned methods (discussed in Sections 2.1.2.1–2.1.2.4) are often combined to achieve enhanced control over the process. Such combinations have been exploited to access morphologies such as the yarn twist (Fig. 2(i)), Janus array fibre films28,35,47 and oriented coaxial fibres.48 Wang et al.36 fabricated highly fluorescent membranes made up of Janus NFs composed of magnetic [Fe3O4/polyvinylpyrrolidone (PVP)] and fluorescent terbium ligand complex Tb(BA)3phen/PVP (where BA is benzoic acid and phen is phenanthroline) NFs. The trifunctional bi-layered composite nanofibrous film was produced using layer-by-layer electrospinning and by systematically altering the process parameters during the electrospinning process.
2.1.2.6 Industrial production. In order to improve production efficiency, and to allow scale-up from the laboratory to industrial manufacture, modifications to the electrospinning equipment are often made. Most commonly the liquid supply device is adjusted commonly to; the use of densely packed needle arrays49 (Fig. 2(j)), separate silk threads used as spinnerets,50 a drum with grooves that rotates in a liquid pool to fill the hole for the supply of liquid (Fig. 2(k)) or a feedstock pool mounted with columns of magnets to directly supply the magnetic liquid51 (Fig. 2(l)).

Generally, the development of electrospinning rigs has been based on the design and combination of the two main components: the solution feeding system and collector. The development of industrial mass production equipment has also contributed to the commercialisation of electrospinning fibres.

2.1.3 Morphology and parameters. The processing parameters that affect the structure and morphology of an electrospun fibre include the applied voltage, liquid flow rate and the distance between nozzle and collector. The applied voltage determines the electric field intensity and charge density of the droplet, which directly influence the diameter of the fibres produced. Generally, a simple increase in voltage, without changing any other parameters, results in a decrease in fibre diameter. However, special cases still exist. For example, beaded fibres can be produced at low polymer solution concentration and at increasing (but low) voltages, the diameter initially increases with decreasing bead density. It is only at significantly higher voltages that a reversal is observed; the fibre diameter decreases and the beads reappear.52 Additionally, a voltage increase can also decrease the uniformity and length of the fibres produced.53 Conversely, increasing the flow rate of the liquid results in an increase in fibre diameter. Sufficient time is required for the solvent to fully evaporate during the fibre forming process, which corresponds directly to the distance between the nozzle and collector. If the distance between the nozzle and collector is too short there is insufficient time for the solvent to evaporate and fused fibres are formed. Similarly, fused fibres are formed when the distance is too large, this is attributed to the reduced electrostatic field strength experienced by the fibres meaning they are not stretched appropriately. In this latter case, the fibres formed often have large diameters which is a result of solvent being trapped within.54

System properties such as polymer molecular weight, concentration, solution viscosity, solvent type and solution electroconductivity also affect the fibre morphology. Electrospinning relies on chain entanglement to produce fibres. The level of chain entanglement is directly related to the solution viscosity, which is intrinsically linked to the polymer molecular weight and sufficiently high solution concentration. Insufficient chain entanglement causes bead-like morphologies instead of continuous fibres. Wang et al.9 studied the effect of co-solvent and polymer concentration on fibre morphology and the phenomenon of microphase separation during solution fibrillation (Fig. 3(m–q)). They found a morphological transition from fibres to beads occurred when increasing the concentration of dimethylformamide (DMF) in the THF/DMF co-solvent system. However, for the polymer system to self-assemble (microphase separate) the quantity of THF present had to be between 65–90 wt%. In another example, Doepke et al.50 investigated nanoparticle concentration when preparing polymer bead/fibre mats for data storage. In this case they found that mechanical dispersion by ultrasonic treatment allowed for higher quantities of nanoparticles to be incorporated in both mats and bead formation without unwanted agglomeration effects. MNPs or their constituent components (e.g. inorganic metal salts, alloys and oxides) are typically added into solution. These materials are highly electroconductive and alter the solution permittivity and conductivity, which in turn affects the creation of the local electrical field, improves the fibre morphology and decreases the fibre diameter.52 Additionally, environmental factors such as humidity and temperature can also influence the fibre structure. Typically, these conditions relate to the speed of solvent volatilisation and thus affect the overall fibre morphology.

In order to obtain the desired morphology or functionality for the target application, the electrospun fibrous membrane is often post treated. The surface of the NFs can be coated with functional entities (e.g. collagen57) or heat treated to chemically crosslink the polymer to improve mechanical strength and/or prevent dissolution (e.g. for temperature-controlled drug release).1 Sandwich structure fibrous membranes have also been prepared by thermally treating the electrospun sample post-deposition.58 The magnetic mat is sandwiched between two non-magnetic mats before an alternating magnetic field is used to induce magnetic heating, which in turn thermally bonds the nanofibrous mats together.

Overall, the development of solution supply systems has provided a rich and varied internal structure of individual fibres and the development of the collector has resulted in a diverse range of inter-fibre structures.

2.2 Magnetic materials and properties

Magnetic nanomaterials such as Fe3O4, α/γ-Fe2O3 (hematite/maghemite) and MFe2O4 (M = metal) have received considerable attention due to their properties and potential applications. Herein, we review the magnetic materials that have been used in electrospun polymer-based NFs and how this relates to their final structure and properties.
2.2.1 Fe3O4. Fe3O4 (magnetite) is one of the most commonly used magnetic materials due to its ease of preparation, biocompatibility, high surface area, catalytic activity, electrical conductivity, low toxicity and almost full spin polarisation at room temperature (RT). Superparamagnetic38,59,60 and ferrimagnetic58,61 Fe3O4 NPs have been prepared with varying saturation magnetisation (Ms) values. The surface of Fe3O4 nanoparticles can also be modified to alter or enhance given properties. Song et al.21 prepared Fe3O4–polyhedral oligomeric silsesquioxane (POSS) particles using a hydrosilylation reaction whilst maintaining an Ms value of the Fe3O4–POSS at 18.77 emu g−1. The POSS was used in the system to improve the stability of the surface potential and charge retention of the mats. Gong et al.62 synthesised Fe3O4-loaded multi-walled carbon nanotubes (MWCNTs) (as a template for the Fe3O4 NPs), achieving Ms as high as 71.549 emu g−1. Fe3O4 NPs can also be mixed with a variety of polymer materials to increase their stability, introduce stimuli-responsiveness and manipulate the interparticle distance and magnetic interactions (Fig. 4). Additionally, polymers can guide the assembly of Fe3O4 NPs to form novel structures.14,33,63–70
image file: d1tc01477c-f4.tif
Fig. 4 Magnetization curves produced when mixing Fe3O4 NPs with different polymeric materials (a) Fe3O4 NPs in PMMA,63–66 (b) Fe3O4 NPs in PAN,33,67 and (c) Fe3O4 NPs in PVP.14,68–70

Since polymeric materials can encapsulate and bind the nanoparticles as a matrix, research groups mix polymer materials with Fe3O4 NPs by uniaxial electrospinning, coaxial electrospinning and parallel-plate electrospinning. The method of electrospinning used affects the structure and properties of the composite material. The Ms of fibres prepared by uniaxial electrospinning increases with increasing mass of Fe3O4.71,72 Savva et al.73 prepared oleic acid-coated magnetite nanoparticles, which show lower saturation magnetisation (∼40 emu g−1) due to the presence of the organic, nonmagnetic oleic acid coating. However, no significant agglomeration phenomena occur during the electrospinning process as exhibited in Fig. 5(a).


image file: d1tc01477c-f5.tif
Fig. 5 (a) TEM bright field image of a PVP/PLLA/OA–Fe3O4 nanocomposite membrane;73 (b) BM image of [Fe3O4/PMMA] coaxial nanobelts;74 (c) BM image of [Fe3O4/PANI/PMMA]//[Tb(BA)3phen/PMMA] Janus nanoribbons;28 (d) FESEM image of PVP NFs;44 (e) TEM image of α-Fe2O3/europium complex [Eu(DBM)3(Bath), where DBM is dibenzoylmethanate and Bath is bathophenanthroline]/PVP composite NFs;44 (f) TEM image of CoFe2O4/yttrium aluminium garnet (YAG):5% Eu3+/PVP composite NFs;75 reproduced from ref. 75 with permission from The Royal Society of Chemistry; (g) TEM image of strontium hexaferrite nanoparticles (SrM-NPs) embedded in a PVA matrix;76 (h) TEM image of SrM-NPs;76 (i) TEM image of NiZn ferrite nanoparticles;77 (j) SEM image of 1% MGNPs-polymer;78 (k) SEM image of 3% MGNPs-polymer;78 and (l) SEM images of 7% MGNPs-PEO.78

Fluorescent magnetic NFs have been targeted in research owing to their suitability in a wide range of applications such as; light-emitting diodes,79 sensors,80 resonators81 and full-colour displays.82 However, heavy losses in fluorescent intensity is observed when Fe3O4 NPs are in direct contact with luminescent compounds.69 In order to circumnavigate this problem, core–shell and Janus structures have been produced as they offer the opportunity to incorporate both components in disparate zones of the material; minimising the direct interactions that would typically occur between them. Shao et al.64 reported the fabrication of tuneable fluorescent colour-electrical-magnetic trifunctional coaxial nanoribbons using coaxial electrospinning. These coaxial nanoribbons exhibited similar magnetic properties (Ms of 18.58 emu g−1) to the corresponding composite nanoribbons (where all components were mixed within the ribbons). Most significantly, the fluorescent intensity and electrical conductivity of the coaxial nanoribbons were considerably higher than those of the composite nanoribbons, demonstrating the importance of architecture derived properties. Fig. 5(b) demonstrates the coaxial nanobelt structure, revealing that the core contains large quantities of dark-coloured Fe3O4 NPs whilst the shell of the coaxial nanobelts appears transparent.74

Another effective method to create the Janus structure is via parallel-plate electrospinning. Gai et al.83,84 prepared Janus nanobelts from Fe3O4/PVP and rare earth complex/PVP which demonstrated desired magnetism–luminescence bifunctionality. The Ms ranged from 3.16 emu g−1 to 10.19 emu g−1 and the results suggest that the magnetism can be tuned via different Fe3O4 NP loadings. The Janus nanobelts exhibited superparamagnetic behaviour using Fe3O4 nanoparticles of approximately 15 nm diameter. When the dimensions of the magnetic component, such as magnetite, drop to less than 20 nanometres, its magnetisation direction can flip randomly under the influence of temperature. However, in this circumstance, magnetite becomes superparamagnetic with only one magnetism domain.85 In another example, Ma et al.86 fabricated Janus NFs with Fe3O4/poly(methyl methacrylate) (PMMA) as the magnetic component and the Ms reached 32.61 emu g−1 when the mass ratio of Fe3O4 to PMMA was 6[thin space (1/6-em)]:[thin space (1/6-em)]1. This is similar to Fe3O4/rare earth complex/PMMA composite nanobelts (32.15 emu g−1) that have also been produced.86 However, the fluorescent intensity of the Janus nanobelts is considerably higher than that of the composite nanobelts. Additionally, luminescent–electrical–magnetic trifunctional materials are also a popular target structure in multifunctional nanocomposites. Lv et al.87 added polyaniline (PANI) to the magnetic half of the Janus structure and the electrical conductivity values of the Janus NFs increased with increasing PANI loading. However, the conductivity of the Janus NFs decreased with increasing amounts of Fe3O4 NPs due to the influence of Fe3O4 on the polymerisation process of aniline. The inner structure of the Janus nanoribbons can be revealed by the transmission light of a biological microscope (BM). As shown in the Fig. 5(c), one side of the Janus nanoribbon contains large quantities of dark coloured PANI and Fe3O4 NPs and, by contrast, the other side is transparent.28

2.2.2 α-Fe2O3 and other iron oxides. α-Fe2O3 is another popular choice of magnetic material due to its high stability, ease of fabrication, appropriate saturation magnetisation and increased acid resistance compared with Fe3O4. Additionally, α-Fe2O3 also has excellent adsorption ability of heavy metal ions and organic pollutants.88 Wang et al.44 fabricated NFs based on α-Fe2O3 nanoparticles and a europium complex to achieve magnetic–photoluminescent bifunctionality. The coercivity of the sample was 327 Oe with a remanent magnetisation (Mr)/Ms ratio of 0.244. Wang et al.44 found that the 5D07F2 transition was higher in the pure complex than in the composites. This was due to the addition of α-Fe2O3 nanoparticles decreasing the symmetry of the coordination environment for the Eu3+ ions. As shown in Fig. 5(d), the pure poly(vinyl alcohol) (PVA) NFs are relatively smooth. After incorporation of europium complexes and α-Fe2O3 nanoparticles into the polymer matrix the average diameter increased; this was due to an increase in viscosity of the feed solution.

Hematite is also often blended with polymeric materials via uniaxial electrospinning. Meng et al.89 produced a paramagnetic nanofibrous composite film with polylactide (PLA), hydroxyapatite and γ-Fe2O3 nanoparticles. The Ms of γ-Fe2O3 NPs was 67.6 emu g−1 whilst the Ms of the film was 0.0492 emu g−1, achieved at a mass ratio of 8.3% γ-Fe2O3 NPs within the film. Alternatively, Khanlou et al.90 prepared γ-Fe2O3 NPs through a chemical co-precipitation process with an Ms of 12.19 emu g−1. Following the chemical co-precipitation, at 5 wt% γ-Fe2O3 NPs, the NPs were added to a PMMA solution. The Ms of the composite produced was then 6.172 emu g−1. Both cases demonstrate that the magnetic properties of MNP blended NFs are not proportional to the mass ratio of MNPs. Polymers do not simply act as a loading matrix but interact with MNPs and mutually influence the overall magnetic properties.

Additionally, there are other iron oxides that can be used to produce magnetic nanomaterials. For example Zhu et al.91 produced core–shell Fe–FeO nanoparticles with an average diameter of 20 nm. The Ms of the Fe@FeO NPs was 108.1 emu g−1 whilst that of the nanocomposite fibres was 30.6 emu g−1; with a nanoparticle loading of 30 wt%. Before electrospinning, the radii of the core and shell was calculated to be 13.2 and 6.8 nm, respectively. However, after electrospinning the radii became 12.7 (core) and 7.3 nm (shell). The shell thickness increase was attributed to an increase in particle oxidation at the extremely high voltages used during the electrospinning process.

Finally, Murillo-OrtÍz et al.76 embedded strontium hexaferrite nanoparticles (SrM-NPs) in PVA NFs. The ratio of Mr/Ms increased by 81 when 30 wt% SrM-NPs were added to the PVA solution. As observed in Fig. 5(g), these nanoparticles have uniform size and have a localised distribution of NPs inside the surface of the NFs. Additionally, they do not show the presence of agglomerates. Fig. 5(h) then shows that the nanoparticles are ordered on the surface of the fibre and aligned with respect to the NFs’ growth. This is a consequence of the nanoparticles’ interaction with the highly intense electric field aligned with the electrodes in a point-plate configuration.

2.2.3 MFe2O4 (M = metal). In typical spinel and inverse spinel structures of magnetite, Fe(II) and Fe(III) ions are distributed in either octahedral or tetrahedral voids. MFe2O4 can be prepared by doping Mn2+, Fe2+, Co2+ and Ni2+ metal ions into the crystal voids of the iron oxide nanoparticles (IONPs).92 This has resulted in nanoscale spinel ferrites MFe2O4 (M = Co, Mn, or Ni) with 1D structures (e.g. fibres) being studied more actively in recent years.93

Chen et al.94 synthesised and modified CoFe2O4 nanoparticles to improve dispersion. The diameter of the CoFe2O4 particles produced was 5 nm and the Ms achieved was 50 emu g−1. The diameter achieved is smaller than that of the bulk materials, due to the size of the CoFe2O4 crystallites and fewer defects being present in the structure. Finally, the CoFe2O4 NPs were mixed with polyacrylonitrile (PAN) and the composite exhibited an Ms of 45 emu g−1. The decrease in Ms is attributed to the non-magnetic material coating (PAN) and its influence on the uniformity and magnitude of magnetisation by extinguishing the surface magnetic moment. Alternatively, Wang et al.95 fabricated Janus NFs using CoFe2O4 to achieve magnetism–luminescence bifunctionality. When the mass ratio of CoFe2O4[thin space (1/6-em)]:[thin space (1/6-em)]PAN was 1[thin space (1/6-em)]:[thin space (1/6-em)]3 the Ms and Hc achieved were 5.09 memu g−1 and 20 kOe, respectively. Additionally, Bi et al.75 electrospun [Fe(NO3)3 + Co(NO3)2]/PVP precursor solution before annealing in air at 700 °C for 4 hours to prepare CoFe2O4 NFs. YAG:5% Eu3+ calcinated NFs were also prepared via the same method. Janus NFs were then fabricated from both the CoFe2O4 NFs/PVP and YAG:5% Eu3+ NFs/PVP solutions, as shown in Fig. 5(f). The Ms of the CoFe2O4 NFs was 41.34 emu g−1 whilst the Ms of the Janus NFs ranged from 3.12–20.32 emu g−1. The observed enhanced performance is attributed to the isolation of YAG:5% Eu3+ luminescent NFs from the CoFe2O4 magnetic NFs. Gonçalves et al.96 prepared composite fibres of CoFe2O4 and poly(vinylidene fluoride) (PVDF). The composites demonstrated an increase in magnetisation with increasing CoFe2O4 content. They also found that the piezoelectric coefficient of the NF composites increased with increasing applied magnetic field. This is a result of the strain-mediated coupling between the magnetostrictive CoFe2O4 nanoparticles and the piezoelectric PVDF matrix. However, when compared with bulk polymers the piezoelectric coefficients were lower. It is speculated that this reduction is due to clamping by the surrounding material; which may significantly reduce the local deformation of the NFs.

Ghanbari et al.97 synthesised CaFe2O4 nanoparticles that exhibit ferrimagnetism before producing cellulose acetate (CA)–Ag–CaFe2O4 nanocomposites by electrospinning. The Ms, of the nanoparticle compared to the NF, decreased from 6.1 to 0.31 emu g−1 whereas the Hc increased from 40 to 78 Oe, respectively. The authors stated that the magnetic moments of the CaFe2O4 nanoparticles are pinned by the polymer chains so that a higher magnetic field is required to align the single domain nanoparticles in the field direction. Additionally, Khan et al.77 prepared Ni0.6Zn0.4Fe2O4 nanoparticles (see the transmission electron microscopy (TEM) image in Fig. 5(i)) with Ms of 26.81 emu g−1. The NPs were then incorporated into composite NFs [with carbon nanotubes and recycled polystyrene (PS)] at 7.5, 15, and 30 wt% to produce fabrics with Ms values of 2, 4, and 8 emu g−1, respectively.

2.2.4 Other magnetic materials. Other active magnetic materials that have also been incorporated into NPs include Ni, Co, and some dopants (such as Sr). Chen et al.98 reported the Ms of pure Ni nanoparticles as 14.3 emu g−1, which was considerably lower than bulk nickel (58.57 emu g−1). The decrease is attributed to the oxidation of the nickel nanoparticles as a consequence of their large surface area. In another study, Gupta et al.99 measured the Ms of pure MnZnFe–Ni as 25.47 emu g−1 before forming a blend with Estane® 5750; a polyester-based segmented polyurethane (PU). The Ms of the blend increased from 1.71 emu g−1 to 6.33 emu g−1, with increasing MnZnFe–Ni content, and the composite NFs formed demonstrated superparamagnetic behaviour. Murillo-Ortíz et al.17 synthesised SrFe12O19 with diameters ranging from 37 nm to 179 nm. The Mr/Ms and Hc of the SrFe12O19 were 0.63 and 6.22 kOe, respectively. The remanent squareness increased by 15% and the coercivity by 1.2% when SrFe12O19 nanoparticles were added to the PVA NFs by electrospinning. Additionally, the thermal stability and arsenic adsorption ability of the NFs was improved upon SrFe12O19 addition.

Erfan et al.78 prepared ferrimagnetic glass ceramics, with a diameter of 10 nm, through the use of high-energy ball milling. The Ms of the magnetic glass ceramic nanoparticles (MGNPs) was 53 emu g−1 and the Hc equal to 88 Oe. The Ms of the composite fibre reached a maximum of 4.16 emu g−1 when the mass ratio of MGNPs was 7%. Low MGNPs concentration (1 wt%) NFs (Fig. 5(j)) appear clear and smooth, however, the roughness and nanoparticle aggregation on the surface of the NF increased at higher MGNP content (Fig. 5(k) (3 wt%) and 5 (l) (5 wt%)).

Min et al.100 fabricated PVA/ferritin superparamagnetic fibres. The interaction between the host PVA hydrogel and the protein shell on the ferritin bio-nanoparticles was controlled by thermal methods to vary the size and concentration of the ferritin clusters. The average size and concentration of the ferritin clusters increased in the PVA NFs when the mixing temperature was raised from 30 to 80 °C. The close proximity of the ferritin cores within the clusters resulted in magnetic ordering and increased magnetisation in some cases.

Additionally, many research groups are now developing novel magnetic nanomaterials. One example is FePt which has demonstrated good chemical stability and high magnetocrystalline anisotropy.101 Another example shows micron size graphene sheets decorated with cobalt NPs to endow magnetism.102 There has also been an increasing focus on producing composite fibres with a range of polymeric materials. However, the mechanism regarding the interaction between the magnetic nanoparticles and polymer materials remains unclear.

3. Electrospun fibrous templates

This section discusses the various approaches used to create magnetic nanofibrous materials by electrospinning template constructs and the post-spinning treatment used: (i) deposition of magnetic NPs; (ii) carbonisation; and (iii) calcination. Fig. 1(b and c) highlights the different routes that have been used in this area of research. Table 2 serves as a comprehensive summary of this section, listing the various systems that have been explored, the magnetic reagent(s), polymer(s), electrospinning and processing approach, the final morphology obtained, major magnetic property and applications (where relevant) in each case.

3.1 Product morphology and corresponding processing

The template method is a cost-effective and scalable route used to produce both pure inorganic magnetic NFs and organic–inorganic hybrid magnetic NFs. Electrospun fibrous templating is also a convenient method for inorganic magnetic nanofibre (MNF) fabrication and the MNFs produced sometimes exhibit an increase in Ms.103 However, the change in Ms depends on a range of complex factors and is not always enhanced. The appearance of a magnetically inert layer in the fibrous structure,104 the formation of ferromagnetic phase105 and smaller grain size are all factors that may contribute to a decrease in Ms in 1D nanostructures prepared by the electrospun fibrous template method.106 The templating procedure of the MNFs (Fig. 1(b and c)) generally encompasses two steps: electrospinning to create a fibrous template followed by post-treatment of the template. A typical template-assisted electrospinning precursor solution is made up of three parts: (i) a magnetic component; (ii) polymer(s); and (iii) a solvent system. The magnetic component is typically in the form of magnetic nanoparticles (MNPs)107 or metal salts. Polymers are added to the system to meet the viscosity demand of electrospinning (i.e. to provide entanglements). The most widely used polymers include PANI,108 poly(ethylene oxide) (PEO),109 polyurethane,107 PVA,110 PAN,111,112 and PVP.48,113–116 The purpose of the solvent is to dissolve the polymer(s), disperse or dissolve the magnetic component used and improve the charge-carrying capacity of solution. Typical solvents used are DMF, dichloromethane (DCM), methyl ethyl ketone (MEK), deionised water, ethanol116 or a combination thereof.107 The viscosity, concentration, surface tension and dielectric properties of the spinning solution exert a great influence on the electrospinnability of the solution and the diameter, morphology, crystallinity and tensile strength of the NFs. Such influences have been referred to in previous studies.117,118 Finally, Fig. 1(c) highlights an alternative route where NFs without any magnetic components are used as templates before MNPs are deposited upon them.

In the sol–gel process, literature reports have demonstrated the use of metal nitrate114,116,119 and acetate114 precursors with polymer(s) in the production of electrospun NFs with uniform distributions, even at very high loadings.120 Although the mechanisms for such processes were not explicitly discussed, the conditions in which the precursor solutions were prepared are reported.48,121,122

Following initial deposition, the obtained fibres are often dried and pre-sintered to evaporate any solvents.94,114–116,119,121,123 This is followed by calcination, at a given heating rate, in order to (partially or completely) remove the polymers to form pure inorganic fibres. Concomitantly, the metal ions are converted to their neutral elemental state48 and the atoms aggregate to form nanoparticles.124 The calcination conditions are adapted to meet the varying demands of a given system. For example, air is an idealised atmosphere if metal oxide NFs125 are targeted, whereas hydrogen (H2),126–128 often in combination with an inert gas such as argon (Ar),129–132 and ammonia gas (NH3) are adopted as a reduction (often referred to as de-oxidation in the literature) source. In some cases, the calcination atmosphere is selected to generate oxygen vacancies or eliminate impure phases.133 For example, α-Fe2O3 NFs obtained in a study by Guo et al.134 were heated in an NH3 atmosphere and the crystallites were transformed into Fe3O4 crystallites. Additionally, when the temperature was raised, Fe2N NFs were obtained. The electrospinning parameters, environmental conditions, physical and electrical properties of the spinning solution, and calcinating conditions (temperature, rate, profile and duration) all enable the crystal structure, morphology and ultimate properties of the final product to be manipulated.135 For example, the preparation of mullite–nickel nanocomposite NFs132 was performed through two heating stages. The first was used to convert Ni2+ to Ni NPs in a reducing atmosphere between 550–750 °C before the mullite phase was formed at 1000 °C. However, if the first stage was not allowed to proceed for sufficient time (to permit Ni2+ to be fully reduced to Ni NPs), an undesirable spinel phase mixed with a mullite phase was formed during the second heating stage. In a similar procedure, a Fe/Ni alloy was prepared from the reduction of NiFe2O4 (300–600 °C) to produce Fe–Ni alloy nanoribbons.126,136 The morphology of the Fe–Ni alloy precursor (body-centred cubic or face-centred cubic) was influenced by the heating temperature used during the reduction stage.

Often the calcinated products are formed as powders and do not maintain their structure.41 Efforts have been made to resolve this problem by shifting towards the production of fibre mats that offer the added benefit of being flexible, easy to produce, recyclable and cost-effective. In one example, flexible fibre mats have been utilised in the development of magnetic devices; offering a new design method for electromagnetic shielding systems.121 Additionally, the ease of recycling the mats and the enhanced structural stability, when compared to powders, is highly desirable in applications such as waste water treatment.118 The key factors that influence the flexibility of these fibrous mats are the uniform distribution of NPs throughout the NFs,110 the porous structure of the mats110 and the interfacial energy between NPs and NFs.110 Consequently, attempts have been made to improve the interfacial interactions between the MNPs and the matrix.94 For example, zein (a maize protein) was used as an adhesive between NiFe2O4 NPs and SiO2 NFs to render the resultant fibre mats with extraordinary flexibility.110 In addition, Wang et al. reported that oxygen plasma treatment was an effective way to keep the membrane integrated during the stage where the organic component is removed.41 Finally, flexible calcinated γ-Fe2O3/C NF mats have also been prepared by combining electrospinning, hydrothermal synthesis and calcination.121,137

NFs with different morphologies can also be fabricated by changing the shape of the liquid supply nozzle or by adjusting the calcination parameters used to remove the polymer construct. In the first stage, when polymers are present, fibres decorated with inorganic particles are obtained. Once the polymers have been removed, necklace like NFs, short NFs, nanobelts, core shell NFs, hollow NFs and regular NFs can be obtained. Fig. 6 shows the various types of MNFs that can be targeted by electrospinning, as discussed in the following sections.


image file: d1tc01477c-f6.tif
Fig. 6 Example SEM images of: (a) fibres with particle deposition;107 (b) necklace-like NFs;138 (c) short NFs;139 (d) nanobelts;111 (e) core–shell NFs;140 reproduced from ref. 140 with permission from The Royal Society of Chemistry; (f) hollow NFs;134 reproduced from ref. 134 with permission from The Royal Society of Chemistry; (g) nanotubes;141 reprinted with permission from ref. 141. Copyright (2021) American Chemical Society; (h) 3D crosslinking NFs;109 and (i) regular NFs (calcinated).142 Specific details: (a) Fe3O4@PU NFs, Fe3O4 NPs at 1 mg ml−1; (b) SrFe12O19 NFs calcinated at 1000 °C; (c) BiFeO3 NFs calcinated at 550 °C; (d) Fe3O4/C NFs calcinated at 800 °C; (e) CoFe2O4–Pb(Zr0.52Ti0.48)O3 NFs calcinated at 750 °C; (f) Fe2N NFs calcinated at 400 °C; (g) SnO2 NFs calcinated at 500 °C; (h) Fe3O4–alginate (SA)/PVA crosslinked NFs; and (i) NiFe2O4/multi-walled carbon nanotube (MWCNT) carbon-based NFs (CNFs) calcinated at 850 °C.
3.1.1 Decorated fibres. Fig. 1(c) shows how fibres decorated with particles can be created and Fig. 6(a) shows an example of such surface topology. Polymers are electrospun to form fibres and then MNPs are coated onto the fibres either by in situ synthesis109,143 or post-spinning deposition.107,108,112 Both approaches face the issue of aggregation, however, there have been few research efforts focused on resolving this problem.107,110,143 In one notable example, in situ synthesis was studied, focussing on the influence of drying mode on the morphology of iron oxides upon polyimide (PI) NFs.143 This study concluded that drying under vacuum was more effective than drying in air to allow for uniform particle distribution. When using the deposition method, specific techniques are used to provide uniform distribution of the MNPs. These techniques include use of a surfactant,110 polyol immersion107 and layer-by-layer assembly.144 In some cases, where the polymer completely remains, cross-linking agents can be used to form 3D cross-linked networks109 to improve water resistance and mechanical integrity.145
3.1.2 Necklace-like nanofibres. In single nozzle electrospinning, where amorphous polymer templates are removed by calcination, NFs will inevitably transform from smooth morphologies to rough surfaces before necklace-like structures are eventually formed (Fig. 6(b)). This happens when the calcination temperature reaches a relatively high value. For example, Xiang et al.122 found that the calcinated Co0.5Ni0.5Fe2O4 NFs formed the necklace-like morphology when the temperature was greater than 800 °C. The transmission electron microscopy (TEM) images collected in this study revealed that the rapid growth of nanocrystals, aggregation of small NPs and resultant formation of larger NPs at high temperatures gave rise to their unique morphology.
3.1.3 Short nanofibres. The length of electrospun NFs can reach several tens to even hundreds of microns,146,147 where ‘short’ NFs are defined as those that are less than ten microns in length (Fig. 6(c)). Short/broken fibres are typically generated under two circumstances: (1) when the NPs experience uneven thermal stress and they aggregate; and (2) by the inhomogeneous distribution of sacrificial templates.139,148 In both cases, the local area will fracture readily. For example, Sakar et al.139 fabricated BiFeO3 NFs at varied spinning voltages (8, 10, 15 and 20 kV) and found that 10 kV or less was not sufficient to homogenously distribute the polymer. The calcinated products then have the tendency to be broken at the area where the polymer is scarce.
3.1.4 Nanobelts and nanoribbons. Necklace-like NFs (Fig. 6(b)), nanobelts (Fig. 6(d)) or nanoribbons can be fabricated with minor adjustment to the composition of the spinning solution and electrospinning experimental conditions.111 One possible formation mechanism suggests that if the solvent evaporates rapidly during jet flow, a columnar transition state structure is formed with a thin elastic polymer skin. This structure buckles gradually and finally becomes belt-like.119,149 Additionally, in this process, specific polymers (such as PMMA) are found to be beneficial towards the formation of the flat shape belt NF morphology.111 Additionally, the effects of calcination have been investigated and show no significant changes to the morphology of the nanobelt following removal of the polymer.
3.1.5 Core–shell nanofibres. An example of the core–shell structure is illustrated in Fig. 6(e). These structures are typically generated by coaxial electrospinning or, alternatively, by using a single nozzle, followed by calcination and finally coating. There have been a number of reports where magnetic core–shell NFs are produced. In one example, Bayat et al. coated polydimethylsiloxane (PDMS) on Fe3O4/C NFs, which were produced via electrospinning followed by pyrolysis at 900 °C, for future use as electromagnetic interference (EMI) shields.150 In a second method, α-Fe2O3 NFs were soaked in a tetraethyl orthosilicate (TEOS) precursor solution whereby a catalysed hydrolysis and condensation polymerisation reaction resulted in a uniform silica (SiO2) coating along the fibres.151 Li et al.3 and Zou et al.152 coated SrFe12O19 and CuFe2O4 NFs with TiO2 and CeO2, respectively, via a soaking and calcination procedure. In addition to these multi-step procedures, as aforementioned, core–shell NFs can be fabricated through a straightforward one-step method by coaxial electrospinning. In this case, two different spinning solutions are loaded into inner and outer syringes. Each jet flow is then formed with a different velocity and collected before being annealed to obtain magnetic–magnetic,153–155 magnetic–luminescent156 or magnetic–electric48,140,157 bifunctional core–shell nanofibrous materials.
3.1.6 Hollow nanofibres. Similar to the construction of core–shell MNFs, hollow fibres (Fig. 6(f)) can be prepared with single158–160 or coaxial nozzles,15,161 both followed by calcination. During the single spinning procedure, both Cheng et al.160 and Chen et al.98 identify the key factor critical in hollow or solid morphology determination. Cheng et al. shows that the gas diffusion rate, from the interspaces between the nanoparticles, is slower than the PVP decomposition. As the polymer decomposition continues, the observed increase in internal pressure forces the nanoparticles to the exterior and hollow fibres form.160 However, the study of Chen et al. highlighted that metallic shell formation is the key factor in hollow morphology determination.103,158 When the NFs are heated in air, metal oxide clusters begin to form on the surface of the fibre until they create a shell. After which, at even higher temperatures, the polymer component degrades to form volatile decomposition products leaving behind the hollow NF structure. This illustrates that hollow structures are difficult to produce in cases where the metal oxide forming temperature is higher than the decomposition temperature of the polymer matrix. Further studies by Zhao et al.159 have determined that a concentration gradient (which causes varied consumption rates of Cu2+ and Fe3+) and difference in gas pressure between the inside and outside of the fibre are also two important factors to consider when preparing hollow CuFe2O4 NFs. On the other hand, Fu et al. determined that phase separation caused by a concentration gradient is the most critical factor in the preparation of hollow MNFs.135 Notably, the organic nanofibrous template can be removed afterwards to create hollow MNFs.162

Nanotubes (Fig. 6(g)) are similar in structure to hollow fibres and the differences between them are subtle and sometimes difficult to identify.163 In some cases, the term ‘nanotube’ is used to describe shorter, defect-free hollow fibres, but there is no clear definition and the term tends to be used interchangeably with hollow NFs across the literature. Accurate heating methods are required for nanotube construction via single nozzle electrospinning and Jiang et al. has reported the successful synthesis of nanotubes several microns in length.164 Additionally, Li et al. obtained NaYF4:Yb/Er/Gd-decorated SiO2 nanotubes via single nozzle electrospinning followed by calcination using a spinning solution comprised of NaYF4:Yb/Er/Gd nanocrystals, TEOS and PVP. The as-spun NFs were annealed between 200–600 °C at a heating rate of 2 °C min−1. In this example, PVP is forced to migrate to the surface once leaving the nozzle resulting in a PVP–silica shell with the magnetic component and TEOS within the core. Upon removal of the organic material, nanotubes were produced.165

In the coaxial spinning procedure, the pure organic part (such as mineral oil, polymers or components that incorporate magnetic ingredients166) can be adopted as the core solution.167 However, the flow rate of the inner solution should be controlled to prevent leakage.166

In addition to the aforementioned traditional single-walled hollow structures, there are novel structures that have been obtained, such as hollow-core–double-shell NFs. Kim et al.158 produced CoFe2O4 core–PANI double shell structures. Initially hollow CoFe2O4 NFs were obtained by calcinating the precursor fibres between 80–550 °C at a heating rate of 5 °C min−1 for 2 hours. The inner and outer surfaces were then coated with PANI via in situ oxidative polymerisation.

3.1.7 Three dimensional networks. Post-treatment can be adopted to enhance the properties of NF mats. If the polymer scaffold is retained in the final product, a cross-linking agent is frequently used to form 3D cross-linked networks109 (Fig. 6(h)) to improve water resistance and enhance mechanical stability.145 For example, Gao et al. adopted N,N′-trimethylene-bis[2-(vinylsulfonyl)acetamide] as a chemical cross-linking agent for PVA. Whilst a slight weight loss (<0.4%) was observed after immersing the mat in hot water for two hours the cross-linked PVA NF mats demonstrated enhanced water resistance.168 Chemically cross-linking the polymer chains prevents dissolution of the macromolecules and changes the mechanical properties of the constructs, due to the creation of a chemically bound mesh that no longer relies on physical entanglements for structural integrity.
3.1.8 Regular nanofibres. If the NFs are calcinated solid fibres that cannot be categorised into any of the aforementioned morphologies, they are defined as regular NFs (Fig. 6(i)). Regular NFs are then subdivided into two categories, depending on whether they have rough or smooth surface morphology. Generally, aggregated crystals consisting of NFs appear after the removal of organic components and result in the surface morphology of the NFs being rough. However, there are two exceptions to this rule. The first case is when carbon-containing backbone polymers are carbonised instead of being completely removed. In this case, MNPs are embedded in the continuous carbon phase.94 The second case is when the calcination atmosphere is controlled. For example, Barakat et al.169 revealed that smooth morphologies of nickel NFs were retained even after heat treatment in argon. However, previous studies where hydrogen has been adopted to reduce the nickel precursor NFs do not produce a smooth morphology. To obtain NFs with a uniform size distribution, the surface tension, viscosity and electronic conductivity can be adjusted through the addition of a low molecular weight agent, if optimum conditions cannot be otherwise identified.168

3.2 Processing parameters and magnetic properties

The most common processing parameter for electrospinning NF templates is calcination. Calcinated NFs still possess superior magnetic performance than their zero-dimensional (0D, i.e. nanoparticles) or bulk counterparts. The most apparent property enhancement is in the coercivity (Hc) due to the anisotropic nature of NFs, as compared to NPs.127 The magnitude of the coercivity is related to the anisotropy of the material, which includes magneto-crystalline anisotropy, shape anisotropy and stress anisotropy. However, the NF structure possesses large shape anisotropy as a result of the strict restriction of magnetic moment along the fibre axis.170 For example, LaFeO3 exhibits antiferromagnetism in the bulk, yet is ferromagnetic when in processed as a one-dimensional (1D) nanostructure with an increased coercivity value of 28[thin space (1/6-em)]000 Oe.171 In addition, 1D hard–soft exchange-coupling nanomaterials offer a unique platform for enhancing the maximum magnetic energy product [(BH)max] due to the reduced self-aggregation when compared to their 0D counterparts.124,154 In another case, Lee et al.154 fabricated core–shell Sm2Co17/FeCo NFs by electrospinning, calcination, calciothermic reduction and electroless plating. Relative to its 0D counterpart, the well-dispersed 1D/core–shell nanostructure demonstrates an enhancement in (BH)max (46% increase) as a result of the dense homogeneous soft magnetic coating as shown in Fig. 7. However, it should be noted that the magnetic properties are not always improved by simply introducing anisotropy (going from nanoparticles (0D) to NFs (1D) nanomaterials). This was highlighted in a study where the Ms value of CoFe2O4 NFs was lower than that of CoFe2O4 NPs and even the bulk counterparts.172
image file: d1tc01477c-f7.tif
Fig. 7 Scheme of expected magnetic performance in a uniform, one-dimensional, hard–soft magnetic core–shell nanocomposite and its counterpart.154 Adapted with permission from ref. 154. Copyright (2021) American Chemical Society.
3.2.1 Morphology. As aforementioned, morphology is another factor that influences the magnetic properties of the final NFs. Manipulating the electrospinning voltage can cause the morphology of the MNFs to go through a series of changes. However, this has only been demonstrated in a limited number of studies. In one example, at lower voltages (∼8 kV) the NFs aggregate and by gradually increasing the voltage pass through the following formations: rods (∼10 kV), fibres (∼15 kV) and finally, belts (∼20 kV). The Ms values are then shown to decrease in the sequence of belts > rods > fibres > aggregated fibres.139 In another notable example, Zhao et al. highlighted that SnO2 nanotubes exhibit stronger room-temperature ferromagnetism than the corresponding NFs. This observation is attributed to a larger number of surface defects that result from the Sn interstitial and O vacancies having lower formation energies than the Sn vacancy and the surface area-to-volume atom quantities for the nanotubes being greater.173 Additionally, the average particle size for the nanotubes (30 nm) was larger than that of the NFs (20 nm). Similarly, the average outer diameter of the nanotubes (110 nm, with a 10 nm wall thickness) was also larger than that of the NFs (100 nm).
3.2.2 Calcination. The magnetic properties of NFs are strongly dependent on calcination. Diamagnetic as-spun NFs have been shown to exhibit ferromagnetic behaviour after calcination. The calcination parameters such as rate, temperature and calcination atmosphere exert influence on the final properties of the NFs.174 When the organic components are completely removed, the calcinated products are typically composed of nanocrystallites of various size. Overall, the magnetic behaviour of the nanocrystallites assembly is determined by the shape and size of the crystalline NPs and the interactions between them.114,120,159,175
3.2.2.1 Calcination temperature. The relationship between calcination temperature and Ms is often system dependent. Generally, for a given system, lower thermal treatment temperatures result in smaller particles being formed (down to a critical minimum size) and the smaller particles exhibit smaller Ms values.106,114,119,122,132,133,170,172,176–188 Larger particles have less magnetic contribution from the surface (whilst the contribution from the interior concomitantly increases) and thus the Ms value increases.114 However, the smaller Ms values observed at low calcination temperatures are not always only due to the decrease in size of the particles. For example, in Liu's study, the CoFe2O4 NFs were calcinated at 700, 750, 800, 850 °C, respectively. In addition to the increase in particle size, the XPS results suggested that as the calcination temperature increases, the migration of cobalt ions to the tetrahedral sites as well as iron ions to the octahedral sites increases, thus also contributing to the larger Ms observed.189 However, there are examples that do deviate from this trend. For example, the Ms values of CuCo2O4 fibres have been found to decrease with increasing calcination temperature.190 In this case, the decrease is attributed to surface distortions brought about by interactions between oxygen atoms and transition metal ions. Ti0.9V0.1O2 NFs were also shown to exhibit a decrease in Ms with increasing heat treatment temperature, which was assigned to the increase of nearest-neighbour V ions with direct antiferromagnetic coupling.191

In the study of Liu et al., the Ms value of α-Fe2O3/TiO2 NFs decreased with increasing calcination temperature. The authors ascribed this to an increased density of α-Fe2O3.177 The decrease in Ms with arising calcination temperature was also found in other studies.192 In more recent work, Ponhan et al.193 determined that the Ms dependency on calcination temperature of ZnFe2O4 NFs exhibits through both of the observed trends. Initially, Ms increases with calcination temperature before then decreasing, with crystallite size increasing typically from 19 to 26 nm.

Similarly, the Hc transition with calcination temperature is also system dependent. In most cases, Hc increases with calcination temperature; including that observed for MgFe2O4 nanotubes,135 CoFe2O4 NFs,194 NiFe2O4 NFs,183 MnFe2O4 NFs,186 Ni/mullite NFs132 and Co0.5Ni0.5Fe2O4 NFs.122 However, there are also reports where the Hc transition does not follow this trend. For example, the Hc value of MnFe2O4/C NFs decreases with increasing calcination temperature185 and is attributed to the MnFe2O4/C NFs exhibiting ferromagnetic properties due to the distribution of cations over tetrahedral and octahedral sites. In the study by Lu et al.,106 the Hc value of CoFe2O4 NFs demonstrated an initial increase followed by a decrease with increasing annealing temperature. The same phenomenon was reported in both SrTiO3/SrFe12O19 NFs181 and yttrium iron garnet NFs,170 attributed to the motion of domains experienced at different particle sizes. Furthermore, the same trend was found in hollow CuFe2O4 NFs.159 In this case, higher calcination temperatures gave rise to larger MNPs and thus a higher field force was required to alter the magnetic moment direction; in turn resulting in higher Hc. However, when the particle size exceeded the critical size for a single-multi domain transition, Hc declined. Findings from the study of CoFe2O4 hollow fibres further support this claim as Hc increases with annealing temperature.160 The Hc value of the CoFe2O4 nanobelts119 achieved a maximum value of 1802 Oe before decreasing with further increases in calcination temperature as a result of the particle size being below the critical domain size. In addition, the absence of domain walls and unique rectangular cross-sectional shape contributed to the high magneto-crystalline properties observed.

The final important magnetic property to be discussed in relation to calcination temperature is remanent magnetisation, Mr. In one example, CoFe2O4 NFs, fabricated by single spinning and calcination, and were shown to exhibit a uniform increase in Mr with increasing calcination temperature. This is commonly attributed to an increase in the particle size.194 The same phenomenon was found for MnFe2O4 NFs,186 SrTiO3/SrFe12O19 NFs,181 MgFe2O4 nanotubes,135 hollow CuFe2O4 NFs159 and hollow CoFe2O4 NFs.160 In one contradictory case, the Mr value of CuCo2O4 fibres190 was found to decrease with increasing calcination temperature, which was suggested to be due to weakening of the superexchange interaction between the Cu3+ and Co2+ cations.

In some fabrication methodologies, the heat treatment encompasses two stages; calcination and reduction. The magnetic properties are known to be dependent on the reduction temperature. For example, the reduction temperature of Fe/Ni alloy nanobelts126 show an increase followed by decreasing trend in Ms, whereas the Hc initially decreases before remaining almost constant with increasing temperature. In this example, the Fe/Ni alloy nanobelts were reduced from NiFe2O4 and the initial relative low temperatures used resulted in the observed partial reduction of Ms.

The relationship between calcination temperature and Ms, Hc and Mr has been identified as significantly system dependent. Factors such as density, surface distortions, cation distribution, particle size and cation interaction have all been shown to contribute to the final magnetic properties of the NFs produced.


3.2.2.2 Calcination atmosphere. The calcination atmosphere used can have a significant effect on the magnetic properties of NFs. For example, SrFe12O19/CoFe2O4 obtained from as-spun NFs calcinated in air showed a Hc value of 3190 Oe. On the other hand, SrFe12O19/FeCo obtained from as-spun NFs calcinated in air and then reduced in hydrogen showed a significantly lower Hc value of 1249 Oe. This change is attributed to the close contact of the soft and hard phases which in turn results in strong exchange coupling. In addition, Fe2O3 hollow NFs obtained from as-spun NFs calcinated in air were superparamagnetic and achieved an Ms value of 4.34 emu g−1, whilst Fe3O4 and Fe2N hollow NFs obtained from as-spun NFs calcinated in air and then reduced in gaseous ammonia achieved Ms values of 82.99 emu g−1 and 2.07 emu g−1, respectively. Finally, the Fe3O4 hollow NFs exhibited ferromagnetism whilst the Fe2N hollow NFs demonstrated superparamagnetism.134,153 The morphology of the fibre can also be altered depending on the calcination atmosphere used and, as previously discussed, is directly related to the magnetic properties of the final fibre. In one example α-Fe2O3 nanotubes and Fe3O4 NFs were produced when the calcination atmospheres were air and argon, respectively. Nanotubes were formed due to the lack of oxygen present as gas diffusion from the decomposed organic component (PEO) drives the nanoparticles from the inside to the outside of the fibres.195
3.2.3 Magnetic component loading. The most influential factor on the overall magnetic performance of NFs are the components used to make them, rather than how they are made. This is particularly evident in cases that involve the combination of magnetic inorganic constituents and non-magnetic polymers. Typically, Ms has a linear correlation with the quantity of magnetic components,107,156,196–199 whilst the addition of non-magnetic components generally impairs the Ms value.200–202 Therefore, the Ms values of NFs where polymers are retained are lower than those observed for fibres that consist of pure MNPs.94,107 Another common example, reflecting this logical observation, is Ms enhancement after carbonisation. Again, due to the loss of the non-magnetic compounds during the procedure. With ascending carbonisation temperature, Ms will increase when the size and coarsening of MNPs within the NFs increases and if a new phase with larger Ms comes into force.198 However, one report claims that the Ms value in their system was independent of shape and size of the MNPs and is only affected by the total quantity of magnetic atoms.198

Moreover, carbonisation often induces the transformation from superparamagnetism to ferromagnetism in electrospun NFs.194 This is due to the formation of a new phase, such as the carbide crystal, which destroys the single crystal domain of the original NFs.172 It has also been claimed that enhancement of particle size leads to higher magnetocrystalline anisotropy.197,198

3.2.4 Types of polymer. The different types of sacrificial polymer result in varied magnetic properties in the final NF. For example, NiO NFs obtained from poly(2-ethyl-2-oxazoline) (PEtOx)/nickel(II) acetate tetrahydrate precursor were shown to exhibit antiferromagnetism, whereas NiO NFs obtained from styrene–acrylonitrile random copolymer (SAN)/nickel(II) acetate tetrahydrate precursor exhibited ferromagnetism at RT.203 The authors suggested that the change in magnetism occurs because of the distinct degradation mechanism of the different polymers. The degradation of SAN begins with elimination of the nitrile group followed by random scission of the backbone. In contrast, PEtOx begins with random chain scission followed by elimination of the side chains. Additionally, the activation energy for grain growth and corresponding grain sizes were also different. Both factors lead to non-stoichiometry of the NiO NFs.
3.2.5 Polymer concentration. Polymer concentration notably influences the electrospinning process, which determines the morphology and magnetic properties of the fibres produced. Typically, increasing the concentration of the polymer produces fibres with increased diameter, uniformity and tensile strength.204 The variation in morphology, particularly NF diameter and uniformity, is also often attributed to the solution properties (polymer and solvent) on the electrospinning process; most significantly the viscosity of the electrospinning solution. The viscosity, alongside the surface tension, ultimately determine the ability of a given solution to be electrospun. Therefore, it is unsurprising to see examples in the literature that describe increased solution viscosity due to higher concentrations of polymer resulting in NFs with increased diameter.205 In one investigation, the effect of PVP concentration on the morphology and magnetic properties of SrFe12O19 nanobelts was investigated.116 It was found that increasing the PVP concentration from 8.5–12.3 wt% led to the increase in the width of the nanobelts produced. Additionally, the diameter of NPs were reduced and the NPs’ overall size was largely below the single domain size. As a direct result of this, Hc achieved a maximum value of 7310 Oe (the highest known value in the literature to the best of our knowledge) of all the pure 1D SrFe12O19 structures. This value is also close to the theoretical limit.116
3.2.6 Dopants. Doping has been shown to affect the crystal size of magnetic materials, which influences the magnetic properties of the final MNFs. The extent of this effect depends on the properties and loading of the dopant.206–212 Typical dopants that have been explored to enhance the magnetic properties and direct the morphology of MNFs include graphene oxide (GO),213 Co,214 Ni,215 Cu210 and Zn.216 For example, GO-doped Ni0.8Gd0.2Fe2O4 NFs showed a transformation from fibre to ribbon morphology and a decrease in Ms with increasing GO content. Additionally, Ni-doped Fe NFs were shown to exhibit lower Ms and Hc values compared to pure Fe NFs.128 Ni-doped CoFe2O4 NFs were also found to show a decreasing trend in Ms and Hc with increasing Ni2+ concentration, which was due to the lower magnetic moment and magnetocrystalline anisotropy of Ni2+ compared to Co2+ ions.215 It is important to note that the change in magnetic properties with dopant concentration is not always uniform209 and can contradict known theories (such as Brown's theory217) where Ms and Hc are inversely related. For example, the Ms value of Zn-doped CuFe2O4 NFs initially increases before decreasing with increasing Zn2+ concentration due to increased exchange interaction between ions within the lattice.216 Co-Doped ZnO NFs were fabricated by electrospinning a PVA/Co-doped ZnO suspension followed by sintering in air.218 Doping ZnO NFs with 1.8–7.2 wt% Co was shown to change the material properties from ferromagnetism to ferrimagnetism, as shown by the temperature-dependent magnetisation curves and Curie–Weiss fits in Fig. 8. The origin of ferrimagnetism in Zn1−xCoxO NFs is attributed to the substitution of Co for Zn and the corresponding change of characteristics in the electronic state.
image file: d1tc01477c-f8.tif
Fig. 8 The effect of doping ZnO NFs with Co2+ on magnetic properties, illustrated through: (A)–(D) temperature-dependent magnetisation (MT) curves of Zn1−xCoxO (x = 0, pure ZnO nanowires) NFs with 1.8, 4.4 and 7.2% Co, respectively. This visually shows the transition from ferromagnetism to ferrimagnetism. (E) Co2+ doping-induced ferromagnetism to ferrimagnetism crossover.218 Reproduced from ref. 218 with permission from The Royal Society of Chemistry.

Doping has also been known to lead to the production of NFs with unique properties. For example, hole-doped manganites, with the general formula R1−xAxMnO3 (where R refers to a trivalent rare earth element and A is a divalent alkaline earth element) were found to exhibit colossal magneto resistance (CMR). La0.33Pr0.34Ca0.33MnO3 NFs (prepared at a calcination temperature of 600 °C) exhibited CMR over a wide temperature range and the magneto resistance reached a maximum value of 95% at the metal–insulator transition temperature of 70 K under 7 Tesla.219 In another case, doping α-Fe2O3 nanotubes with V2O5 was found to increase Ms, and decrease Hc, which was attributed to the aforementioned Brown's theory; where Hc and Ms are inversely proportional to one another.217 In other work, after Co doping, SrTiO3 was endowed with room temperature ferromagnetism due to Co addition and oxygen vacancies,130,173 which has also been demonstrated in the study of SnO2 nanotubes.220 However, Mohanapriya et al. suggests that ferromagnetism in Mn-doped SnO2 NFs is induced only by precipitated impurity phases instead of pure SnO2 or dopant.221 Nevertheless, Mn doping was found to impair the ferromagnetism of BaTiO3 and resulted in dia- and paramagnetic behaviour of the BaTi0.9Mn0.1O3 NFs that were produced.222 Fe-Doped NiO NFs were prepared as a diluted magnetic semiconductor (DMS) and, notably, doping was shown not to affect the fibre diameter or surface morphology.223 Finally, the addition of GO dopant to CoFe2O4 NFs demonstrated enhanced crystallinity, which boosted Ms values at low loadings, but was then shown to exhibit a slight decrease in Ms with higher GO loadings.224

3.2.7 Alloys. Alloy systems with nanoscale dimensions often demonstrate outstanding magnetic properties. There have been a number of reports where alloys have been processed into NFs to create magnetic materials. For example, Zhang et al.128 reported Fe–Ni NFs for the first time and demonstrated that the use of an Fe–Ni alloy resulted in superior ferromagnetic properties. Additionally, it was suggested by Jiang et al.126 that an Fe–Ni alloy also produces NFs with high saturation magnetisation, permeability, Curie temperature (Tc), low coercivity and low energy loss. FeCo is another promising soft magnetic nanomaterial which exhibits low magneto-striction, high saturation magnetisation, high resistivity, small coercive forces, high Curie temperature and high magnetic anisotropy energies.225 Finally, a range of novel Sm2Co17 NFs were fabricated by Lee et al. reporting a decrease in Ms and concomitant Hc increase with arising Sm content within the alloy.206

4. Applications

In this section, the application of magnetic materials, and how this relates to their final properties, are explored. This aptly highlights the versatility of electrospun magnetic materials whilst concomitantly highlighting the need for future work into their production and design.

4.1 Electromagnetic shielding/absorption

With the extensive utilisation of electronic devices for civil and military purposes there is a strong desire to reduce electromagnetic (EM) pollution due to its associated negative impact on human health, national defence security and electronic safety.121,226–231 EM wave shielding, or absorption materials, are designed to address this challenge. Fig. 9 shows a schematic diagram that illustrates the working mechanism of so-called microwave absorption materials (MAMs). When incident waves reach the surface of the material, they are either reflected off the sample or transmitted into a porous channel; where multiple reflections occur and the energy is dissipated. The difference between MAMs and electromagnetic-induced (EMI) shielding materials is that MAMs require EM waves to be transmitted into the materials to be dissipated rather than simply reflected at the surface.150,196 MAMs achieve EM attenuation through a combination of magnetic loss, electric loss and geometry (morphology).226,228 Magnetic loss originates from the eddy current effect, natural ferromagnetic resonance and exchange resonance. For ferromagnetic materials, the magnetic loss ability is related to initial permeability (μi), as shown in eqn (1), where a and b represent constants relating to the material being used, and λ, k, ξ, Ms and Hc are the wavelength, magnetisation constant, elastic strain parameter of a crystal, saturation magnetisation and coercivity, respectively. Eqn (1) shows that high Ms and low Hc values are required to increase μi, to deliver better magnetic loss ability.226,232
 
image file: d1tc01477c-t1.tif(1)
For electromagnetic wave absorption (EWA) properties, complex permittivity εr = ε′ − ′′ and complex permeability μr = μ′ − ′′ are two critical factors in microwave absorption property determination. The real part (ε′ or μ′) corresponds to the energy storage whilst the imaginary part (ε′′ or μ′′) corresponds to the energy dissipation. Dielectric loss and magnetic loss are expressed by image file: d1tc01477c-t2.tif and image file: d1tc01477c-t3.tif, respectively.196,226,228

image file: d1tc01477c-f9.tif
Fig. 9 Electromagnetic wave attenuation through a shielding material.233

Reflection loss (RL) is calculated from eqn (2)–(4):196,226,228

 
image file: d1tc01477c-t4.tif(2)
 
image file: d1tc01477c-t5.tif(3)
 
image file: d1tc01477c-t6.tif(4)
where, Zin and Z0 are the input impedance of the absorber and impedance of free space, ε0 and μ0 are vacuum permittivity and permeability, and εr, μr, f, c and t are the relative complex permittivity, permeability, frequency of the electromagnetic wave, velocity of electromagnetic waves in free space and thickness of the absorber, respectively.226 Accordingly, thickness plays an important role in the magnitude of RL.226,227 RL decreases and the absorption peaks shift to lower frequencies with increasing absorber thickness. Dielectric and magnetic losses also influence RL and are used in combination to improve the microwave absorption performance.228 Additionally, crosslinking the structure can cause a geometric effect that contributes to RL, as demonstrated with Fe/C cross-linked network structures.228

Carbon fibres can be equipped with magnetic properties through the addition of MNPs to create EM absorbing materials. The ultimate EM wave absorption capability is a direct result of both the dielectric and magnetic properties, with higher Ms and lower Hc values also contributing to the magnetic wave absorption properties.226,228 Notably, NiFe2O4/multi-walled carbon nanotubes (MWCNTs)/carbon nanofibrous membranes, with thicknesses of 2–5 mm, have been created as MAMs that achieve high permeability and reflection losses >20 dB (5.36–18 GHz).142

4.2 Separation: pollutant treatment and catalysis

Functional pollutant absorbers and photocatalysts are used extensively in wastewater treatment. However, the challenge on how to avoid secondary pollution of any residual absorber or catalysts remains a significant problem. One method used to address this issue is the incorporation of electrospun NFs with unique magnetic properties. After degradation, these materials are easily recycled by use of a permanent magnet.45,158,166,177,199,200,202,234–247 For example, Liang et al.248 prepared multifunctional switchable chemosensors for Hg2+ with fluorescent probe 1-benzoyl-3-[2-(2-allyl-1,3-dioxo-2,3-dihydro-1H-benzo[de]isoquinolin-6-ylamino)ethyl]-thiourea (BNPTU), PNIPAM copolymer moiety for thermo-responsiveness and on–off photoluminescence and Fe3O4 NPs for easy removal. Chang et al. deposited bismuth oxyiodide nanoflakes on α-Fe2O3 NFs via the SILAR (successive ion layer absorption and reaction) method.237 The final product reached a maximum specific surface area of 35.27 m2 g−1, rhodamine B degradation efficiency of 98% after 2 hours of visible light irradiation and an Ms value of 0.9 emu g−1, demonstrating its potential for use as practical separable photocatalysts. As these MNPs render the fibres ferromagnetic, or superparamagnetic, they can be attracted to external magnetic fields.239,249–251 In addition to this, it has been shown that under stronger magnetic fields the catalysts can be demulsified and therefore recycled back for reuse.252

4.3 Tissue engineering scaffolds, cell culture and differentiation

Electrospun NFs have been extensively explored as scaffolds for regenerative medicine. Recently, superparamagnetic fibrous scaffolds have been discovered to have a positive effect regarding cell differentiation (Fig. 10). Furthermore, after implanting the scaffold into the human body, the magnetic property enables the device to be manipulated remotely. The most common approach in this area is to combine MNPs, such as γ-Fe2O3,253–255 Fe3O4,8,85,256–259 CoFe2O4 and bioglass,260 with bio-compatible polymers [such as PVA,253–255,260 polycaprolactone (PCL), poly(L-lactide), poly(lactic-co-glycolic acid) (PLGA), poly(succinimide) and poly(aspartic acid)8,85,256–259] to create flexible composite fibrous mats. These mats have been shown to exhibit enhanced mechanical properties compared to pristine non-magnetic NF mats. NFs with a dense, uniform, surface coverage of MNPs are targeted, however, their production remains a challenge.144,242
image file: d1tc01477c-f10.tif
Fig. 10 Schematic showing the mechanism and effect of the IONP-assembled electrospun scaffold on the cells.144 Adapted with permission from ref. 144. Copyright (2021) American Chemical Society.

There have been attempts to employ electrospun fibrous templates to create tissue engineering scaffolds, where such methods appear to facilitate better distribution of MNPs on the surface of the NFs. For example, the layer-by-layer assembly technique was used to afford virtually continuous, compact and uniform α-Fe2O3 nanoparticle immobilisation on PLGA/PCL electrospun scaffold surfaces.144 In this example, the superparamagnetic scaffold was found to significantly enhance the differentiation of adipose-derived stem cells for the treatment of osteogenesis. Moreover, the capping layer brings about auxiliary benefits such as hydrophilicity, elasticity of the interface and the affinity for stem cells. In other work, composite Fe3O4/silk fibroin NFs were compared to silk fibroin fibrous templates with coated Fe3O4 NPs to show the effect of the fabrication method on the final properties of the fabrics. The composite material (electrospun from a single solution) was more effective as a cell scaffold whilst the fibrous mat coated with MNPs was more suitable as a magnetic-sensitive interface.8

4.4 Hyperthermia treatment

The treatment of cancer cells includes chemotherapy, radiation therapy and surgery. Whilst these treatments have been used for decades it has been suggested that methods typically used to treat hypothermia could be an effective and ideal approach. This is based on the incorporation of superparamagnetic NFs which can either give out a fatal amount of heat to kill the cancer cells or the raised temperature of targeted tissues can enhance the effects of other therapy (such as drug release as shown in Fig. 11) used to destroy the cancer.107,109,145,261,262 Superparamagnetic fibres produce a thermal effect in response to an applied alternating magnetic field. The significant advantage of superparamagnetic NFs is their demagnetisation in the absence of an external field. Non-toxic, biodegradable polymeric matrices109 serve as carriers to provide mechanical support alongside other performance-enhancing properties, such as inhibiting the overgrowth of malignant tissues.107 The further merit of NFs in biomedical applications lies with their large surface area-to-volume ratios that lead to faster degradation of the scaffold after they have served their purpose. A large ratio of superparamagnetic NPs relative to polymer is required to provide an adequate number of accessible sites and immobilise the NPs on the NFs.145 There are reports where the number of MNPs on fibres have been quantified. Chen et al. used Fe2+ to chelate with alginates before fabricating a crosslinked Fe3O4–SA/PVA mat by chemical co-precipitation. Immobilising the NPs on the fibre was then shown to eliminate cytotoxicity effects towards human lung fibroblast cells.107,109
image file: d1tc01477c-f11.tif
Fig. 11 Design concept for a smart hyperthermia NF system that uses MNPs dispersed in temperature-responsive polymers.1

4.5 Drug delivery

Recent developments in nanotechnology have led to polymeric matrices (e.g. NFs) being exploited as controllable drug release media.263 Encapsulating bioactive agents in a polymer matrix is an effective method for preventing drug degradation in potentially hostile environments (e.g. high or low pH). Electrospun ultrathin fibres exhibit ideal characteristics such as high porosity, large surface area and diverse controllable morphologies to facilitate the development of electrospun fibrous drug delivery systems.88,264

For example, a core–shell fibrous drug delivery system was produced by coaxial electrospinning using Eudragit S100 as the shell and a core composed of PEO loaded with Gd(DTPA) (gadolinium diethylenetriamine pentaacetate hydrate, magnetic resonance contrast agent) and indomethacin (model therapeutic agent).265 Eudragit is insoluble in acidic environments and when used as oral medication protects the core ingredient as it passes through the stomach. The core materials are only released when the fibres reach the intestinal fluids, enabling this core–shell fibre delivery system to achieve targeted drug delivery to the colon. In another example, PVA/ferritin NF hydrogels with controllable magnetic properties were fabricated by partially unfolding the ferritin protein shell at varied mixing temperatures.100 The negative image contrast generated by ferritin in the PVA matrix under MRI provided a method for in vivo imaging of the tissue-engineered scaffolds.

Meanwhile, magnetic reagents loaded into electrospun fibres have been shown to demonstrate synergistic effects. Sasikala et al.266 designed and synthesised an implantable magnetic nanofibrous device for hyperthermia treatment comprised of iron oxide nanoparticles and tumour-triggered controlled drug release (bortezomib). The fibres exhibited a synergistic anticancer effect by applying the hyperthermia treatment and drug delivery simultaneously. In other work, a similar design was employed by embedding MNPs (as heat generators) and DOX (doxorubicin, anticancer drug) inside a temperature responsive copolymer of poly(N-isopropylacrylamide)-co-(N-hydroxymethylacrylamide) [P(NIPAM-co-HMAAm)].1 When placed in an alternating magnetic field, the crosslinked P(NIPAM-co-HMAAm) NF mesh showed reversible changes in swelling, which created an ‘on–off’ response allowing DOX to be released.

Beyond their use in externally triggered controlled release, the addition of MNPs can have a varied effect on drug release rate, depending on how the MNPs impact on the system within which they are placed. Demir et al.267 found that RhodB (hydrophilic dye) loaded in PCL NFs was released faster when MNPs were present due to the magnetic interaction between the nanoparticles and drug. Similarly, Haroosh et al.268 also found that the addition of MNPs increased drug release rate. In this case, the inclusion of the MNPs increased the conductivity of the electrospinning solution and also decreased its viscosity, which led to the production of thinner fibres. The larger specific surface area of the thinner fibres resulted in the increased drug release rate observed. Conversely, Wang et al.18 found that the release rate of indomethacin and aspirin (model drugs) was not affected by the incorporation of Fe3O4 NPs within the cellulose matrix; even when the MNPs occupied nearly 50% of the fibre mass.

In short, MNPs can be incorporated into NFs for controlled, triggered drug delivery, or to enhance the properties of the nanofibrous drug delivery system. In the latter case, caution must be made when designing and fabricating the device, as the incorporation of the magnetic nanoparticles can have variable effects on the drug delivery performance of the system.

4.6 Nanogenerators

The development of wearable electronic devices has created a demand for flexible, lightweight, self-powering and energy scavenging technologies.269,270 In 2001, Glynne-Jones et al.271 introduced the piezoelectric vibration-powered microgenerator before various materials, such as organic materials, metals, textiles and papers were developed for use in nanogenerators.234,272–276

Fabricating PVDF fibrous nanogenerators by electrospinning has been widely researched and is considered a good method to enhance energy generation performance.277–281 Im et al.282 fabricated Fe3O4/PVDF composite NFs for use in a triboelectric nanogenerator device. Incorporating PVDF resulted in an increase in the surface area and preferential formation of the PVDF polar β-phase, which, in turn, enhanced the triboelectric performance of the device. Additionally, increasing the Fe3O4 content increased the output voltage initially from 124 to 138 V and enhanced the EMI shielding performance when added in small quantities. However, a decrease in the output voltage (94 V) was then observed due to aggregation when the α-Fe2O3 loading was too high (28.3 wt%). Similarly, the tensile strength of the NF initially increased before decreasing, this time due to the dispersion strengthening mechanism. In another example, Wu et al.283 reported a method to synthesise a lead zirconate titanate textile (using PVP as a scaffold) in which the nanowires were parallel to one another. This nanogenerator generated a 6 V output voltage, 45 nA output current and qualified to power common liquid crystal displays and UV sensors. Additionally, the nanowires were soft and flexible, ideal for use in wearable nanogenerators.

In summary, organic materials have been investigated as potentially useful counterparts in magnetic NFs for use as nanogenerators. As demonstrated, incorporating polymers, such as PVDF, into the NFs has resulted in enhancement to the triboelectric performance of such devices. Further advances have also been made to produce soft and flexible materials to be used in wearable nanogenerator devices.

4.7 Data storage and transfer

It is well known that modern data storage media function by storing information as binary data (marked as 0 and 1). Magnetic materials are widely used for data storage as they can be magnetised into two opposing directions, which can respectively represent the binary codes, 0 and 1. Both simulations and experiments have shown that nanocylinders or beads that are several hundred nanometres in diameter can form a vortex state under small or vanishing external magnetic fields.284 The magnetisation of these nanostructures rotates in a closed loop and is often referred to as a vortex-core. In a fibre, the magnetisation state of the beads controls the signal to transfer from one side of the beads to another. The vortex-core can be switched between the two orientations using short magnetic field pulses,285–287 which can be used to write binary data.

Electrospinning is considered a useful tool to create combinations of nanofibrous mats with embedded beads. The beaded fibres can be fabricated when the polymer solid content in the spinning solution is reduced. Döpke et al.50 electrospun PAN/Fe3O4/α-Fe2O3/NiO beaded fibres from 14 wt% polymer solution and computer simulated signal transfer in the beaded fibres. Without an applied static magnetic field the signal transferred through the beads in the shape of a snake-like gyrotropic precess from one side to another. Upon applying a static magnetic field, the magnetisation of the bead fully oriented along the direction perpendicular to the fibre and the signal was blocked. This ‘on’/‘off’ state could not only be used for data storage, but also as logic elements (such as AND or NAND) for neuromorphic computing. Blachowicz et al.288 further simulated the magnetisation reversal mechanisms under different local spatial distributions and mutual influences of neighbouring magnetic fibres. They found a tendency towards larger coercive fields as the NFs were distributed at a larger random angle range. Fibre mats consisting of two types of NFs (with and without branches) were also simulated. In this case the magnetisation reversal was found to start at smaller negative magnetic fields and end at larger negative fields compared to that of the single nanofibre.

In general, using electrospun magnetic fibres for data storage and neuromorphic computing is still in its infancy but demonstrates promising results. Of the work currently completed there has been a focus on simulating magnetic properties with respect to complex intra- and inter-fibre electrospun MNFs. Blachowicz et al.288 studied the influence of numbers and dimensions of contact points of electrospun NFs, and proposed a scheme to verify the mechanism. However, this application requires further investigation to enable widespread implementation.

5. Summary and future prospects

Current production methods used to make electrospun MNFs can be divided into two categories. The first introduces magnetic components into the electrospinning precursor solutions to fabricate continuous MNFs. The second method uses electrospun MNFs as templates in order to process them into various morphologies and structures. Despite the fact that the magnetic properties of MNPs have been widely studied, the overall magnetic properties of resultant MNFs cannot be predicted since the interactions between polymers and MNPs are not well-understood and are therefore rarely discussed. Many types of MNPs have been added to polymeric materials. In some cases, the MNPs and polymers were simply blended and there were no interactions between them, resulting in a linear trend of Ms with MNP content. However, for many composite systems a non-linear trend is observed. Ms values were often increased, and yet some were attenuated, compared to the simple blended systems. However, to the best of our knowledge, there is no evidence provided to establish the relevant interactions between MNPs and polymer materials and the mechanism behind the non-linear relationship remains unclear. Theoretical modelling is required to predict the relationship between structure and magnetic performance for pure inorganic MNFs. Given that this field is very much in its infancy, novel composite and inorganic magnetic systems have been typically investigated to showcase their unique magnetic performance and are not employed in a given application.

Where specific applications are described there are often challenges that remain to be addressed. Loading magnetic resonance imaging contrast agents via electrospinning was a common method to boost the MRI signal of fibre-based materials for in vivo applications. However, passively imaging the materials is not enough for real medical treatment. Smart or environmentally responsive magnetic fibres are needed not only for location and shape information of the fibres under MRI, but also for linking physiological indices, microenvironment parameters (pH, temperature), degrees of degradation or concentration of bio-factors with MRI signal intensity to provide greater internal body information.

Nanofibres with deposited particles have been found to be useful in wastewater treatment due to their high surface area (as compared with MNFs that have MNPs embedded in the matrix), leading to more catalytically active sites. However, MNPs that have been deposited onto NFs (post-electrospinning) are often detached from the fibrous matrix after several absorption–desorption test cycles, which hampers their viability for long-term use. Therefore, further investigation is needed to identify how MNPs could be adhered to the NFs (to survive hundreds of cyclic tests) would be meaningful work. Sensing technologies for heavy metal ions and toxic gases that can cause ill effect to humans have been summarized elsewhere.251 Integrating existing pollutant removal technologies with advanced nanofibrous systems, alongside in-depth studies on these toxicants would significantly extend their applications.

In summary, advances in electrospinning have allowed a wide range of magnetic nanofibrous materials to be created over the past two decades. These advanced materials have shown huge potential and are set to play a role in smart technologies of the future. To unlock their true potential, significant work is needed to understand the key interactions between the various components in the system to overcome the current drawbacks that have been encountered.

The processing methods of electrospun magnetic composite NFs and pure inorganic NFs from templates, their properties and related applications have been summarised and discussed throughout this review. Alongside this, key areas for future research have been highlighted with the aim of stimulating advances in the development of electrospun magnetic nanomaterials for a wide range of applications.

Glossary

ARAcid red 27
BA-aBisphenol-A, paraformaldehyde and aniline monomer hybrid
(BH)maxMaximum magnetic energy product
BMBiological microscopy
BNPTU1-Benzoyl-3-[2-(2-allyl-1,3-dioxo-2,3-dihydro-1H-benzo[de]isoquinolin-6-ylamino)ethyl]-thiourea
CACellulose acetate
CMCCarboxymethyl-cellulose
CMRColossal magneto resistance
CNFsCarbon-based nanofibers
CSChitosan
CTMBCellulose tris-(4-methylbenzoate)
DCMDichloromethane
DMA N,N-Dimethylacrylamide
DMFDimethylformamide
DMSDiluted magnetic semiconductor
DNA–CTMADeoxyribonucleic acid–cetyltrimethylammonium chloride
EMIElectromagnetic interference
EMSElectromagnetic interference shielding
EWAElectromagnetic wave absorption
α-Fe2O3Hematite
γ-Fe2O3Maghemite
FESEMField emission scanning electron microscopy
Gd(DTPA)Gd(III) (diethylenetriamine pentaacetate hydrate)
GOGraphene oxide
H c Coercivity
HPMCPDehydroxypropyl methyl cellulose phthalate
H s Saturation field
IONPsIron oxide nanoparticles
MADOP(MMA-co-DMA)
MAMsMicrowave absorption materials
MEKMethyl ethyl ketone
MGNPsMagnetic glass ceramic nanoparticles
MMAMethyl methacrylate
MNF(s)Magnetic nanofibre(s)
MNP(s)Magnetic nanoparticle(s)
M r Remanent magnetisation
M s Saturation magnetisation
MWCNTsMulti-walled carbon nanotubes
NFsNanofibers
NPsNanoparticles
P(AN-co-AA)Poly(acrylonitrile-co-acrylic acid)
P123Pluronic
PA6Polyamide-6
PAAPolyamic acid
PAAmPolyacrylamide
PANPolyacrylonitrile
PANIPolyaniline
PBTPoly(butylene terephthalate)
PBZPolybenzoxazine
PCLPolycaprolactone
PDAPolydopamine
PDLLAPoly(D,L-lactide)
PDMSPolydimethylsiloxane
PEIPolyethyleneimine
PEK-CPhenolphthalein polyetherketone
PEOPoly(ethylene oxide)
PETPoly(ethylene terephthalate)
PEtOxPoly(2-ethyl-2-oxazoline)
PF–NaPolyfluorene–Na
PHBPoly(3-hydroxybutyrate)
PHEMAPoly(2-hydroxyethyl methacrylate)
PHVBPoly(3-hydroxybutyrate-co-3-hydroxyvalerate)
PIPolyimide
PLGAPoly(lactic-co-glycolic acid)
PMMAPoly(methyl methacrylate)
PNIPAMPoly(N-isopropylacrylamide)
P(NIPAM-co-HMAAm)Poly(N-isopropylacrylamide)-co-(N-hydroxymethylacrylamide)
POSSPolyhedral oligomeric silsesquioxane
PSPolystyrene
PS-b-PIPoly(styrene-block-isoprene)
PUPolyurethane
PVAPoly(vinyl alcohol)
PVAcPoly(vinyl acetate)
PVDFPoly(vinylidene fluoride)
PVDF-TrFEPoly(vinylidene fluoride-trifluoroethylene)
PVPPolyvinylpyrrolidone
RLReflection loss
RTRoom temperature
SANStyrene–acrylonitrile random copolymer
SEMScanning electron microscopy
SILARSuccessive ion layer absorption and reaction technique
T B Blocking temperature
T c Curie (transition) temperature
TEMTransmission electron microscopy
TEOSTetraethyl orthosilicate
THFTetrahydrofuran
TPEEThermoplastic ester elastomer

Conflicts of interest

The authors declare no competing financial interest.

Acknowledgements

The authors thank the financial support from the National Key R&D Program of China (No. 2017YFC11050003), National Natural Science Foundation of China (No. 51890871 and 21807046), Science and Technology Program of Guangzhou (201907010032), Guangdong Project (2016ZT06C322), National Natural Science Foundation of Guangdong (No. 2018A030310628, 2020A151501744), and the Fundamental Research Funds for the Central Universities (2020ZYGXZR064). HJH thanks W. Joe Homer for support.

References

  1. Y.-J. Kim, M. Ebara and T. Aoyagi, Adv. Funct. Mater., 2013, 23, 5753–5761 CrossRef.
  2. O. A. Inozemtseva, Y. E. Salkovskiy, A. N. Severyukhina, I. V. Vidyasheva, N. V. Petrova, H. A. Metwally, I. Y. Stetciura and D. A. Gorin, Russ. Chem. Rev., 2015, 84, 251–274 CrossRef.
  3. C.-J. Li, J.-N. Wang, X.-Y. Li and L.-L. Zhang, J. Mater. Sci., 2011, 46, 2058–2063 CrossRef.
  4. F. Ahmadi, Y. Sozer, M. J. Donahue and I. Tsukerman, IET Electr. Power Appl., 2020, 14, 282–290 CrossRef.
  5. D. Li and Y. Xia, Adv. Mater., 2004, 16, 1151–1170 CrossRef CAS.
  6. B. Sun, Y. Z. Long, H. D. Zhang, M. M. Li, J. L. Duvail, X. Y. Jiang and H. L. Yin, Prog. Polym. Sci., 2014, 39, 862–890 CrossRef.
  7. Y.-Z. Long, M. Yu, B. Sun, C.-Z. Gu and Z. Fan, Chem. Soc. Rev., 2012, 41, 4560–4580 RSC.
  8. O. Lalegul-Ulker, M. T. Vurat, A. E. Elcin and Y. M. Elcin, J. Appl. Polym. Sci., 2019, 136, 48040 CrossRef.
  9. L. Wang, P. D. Topham, O. O. Mykhaylyk, H. Yu, A. J. Ryan, J. P. A. Fairclough and W. Bras, Macromol. Rapid Commun., 2015, 36, 1437–1443 CrossRef PubMed.
  10. K. Suneet, S. Sridhar, P. Agiwal, M. S. Sridhar, K. Sanyal and S. Jain, Int. J. Hyperthermia, 2019, 36, 545–553 CrossRef PubMed.
  11. Y. Song, L. Xu and J. Sui, Therm. Sci., 2019, 23, 2365–2372 CrossRef.
  12. F. Mohammadkhani, M. Montazer and M. Latifi, J. Text. Inst., 2019, 110, 989–999 CrossRef.
  13. L. Mei, H. Chen, Y. Shao, J. Wang and Y. Liu, High Perform. Polym., 2019, 31, 230–237 CrossRef.
  14. X. Zhou, Q. Ma, W. Yu, T. Wang, X. Dong, J. Wang and G. Liu, J. Mater. Sci., 2015, 50, 7884–7895 CrossRef.
  15. S. Niu, L. Zhang, N. Wang, J. Zhu, W. Zhang, Z. Cheng and X. Zhu, React. Funct. Polym., 2013, 73, 1447–1454 CrossRef.
  16. M. Yang, S. Sheng, Q. Ma, N. Lv, W. Yu, J. Wang, X. Dong and G. Liu, Mater. Res.-Ibero-Am. J. Mater., 2016, 19, 308–313 Search PubMed.
  17. R. Murillo-Ortiz, M. Mirabal-Garcia, J. J. Cruz-Rivera, D. Valdez-Perez, J. R. Martinez, F. Perez-Moreno and A. Lobo-Guerrero, Mater. Chem. Phys., 2019, 234, 151–157 CrossRef.
  18. L. Wang, M. Wang, P. D. Topham and Y. Huang, RSC Adv., 2012, 2, 2433–2438 RSC.
  19. S. Wang, Q. Liu, Y. Zhang, S. Wang, Y. Li, Q. Yang and Y. Song, Appl. Surf. Sci., 2013, 279, 150–158 CrossRef CAS.
  20. S. Wang, Z. Sun, E. Yan, J. Yuan, Y. Gao, Y. Bai, Y. Chen, C. Wang, Y. Zheng and T. Jing, Mater. Sci. Eng., B, 2014, 190, 126–132 CrossRef CAS.
  21. X. Song, G. Cheng, B. Cheng and J. Xing, J. Mater. Res., 2016, 31, 2662–2671 CrossRef CAS.
  22. E. Yan, M. Cao, J. Jiang, J. Gao, C. Jiang, X. Ba, X. Yang and D. Zhang, Solid State Sci., 2017, 72, 94–102 CrossRef CAS.
  23. A. Marmur, Langmuir, 2003, 19, 8343–8348 CrossRef CAS.
  24. L. Wang, P. D. Topham, O. O. Mykhaylyk, J. R. Howse, W. Bras, R. A. L. Jones and A. J. Ryan, Adv. Mater., 2007, 19, 3544–3548 CrossRef CAS.
  25. J. Lin, F. Tian, Y. Shang, F. Wang, B. Ding and J. Yu, Nanoscale, 2012, 4, 5316–5320 RSC.
  26. W. Yu, Q. Ma, X. Li, X. Dong, J. Wang and G. Liu, Mater. Lett., 2014, 120, 126–129 CrossRef CAS.
  27. L. Wang, X. Dong, G. Gai, L. Zhao, S. Xu and X. Xiao, J. Nanopart. Res., 2015, 17, 91 CrossRef.
  28. Q. Ma, J. Wang, X. Dong, W. Yu and G. Liu, Adv. Funct. Mater., 2015, 25, 2436–2443 CrossRef CAS.
  29. P. Jing, J. Du, J. Wang, W. Lan, L. Pan, J. Li, J. Wei, D. Cao, X. Zhang, C. Zhao and Q. Liu, Nanoscale, 2015, 7, 14738–14746 RSC.
  30. J. Tian, Q. Ma, X. Dong, W. Yu, M. Yang, Y. Yang, J. Wang and G. Liu, RSC Adv., 2016, 6, 36180–36191 RSC.
  31. R. Patwa, N. Soundararajan, N. Mulchandani, S. M. Bhasney, M. Shah, S. Kumar, A. Kumar and V. Katiyar, Biopolymers, 2018, 109, e23231 CrossRef.
  32. H. Hu, W. Jiang, F. Lan, X. Zeng, S. Ma, Y. Wu and Z. Gu, RSC Adv., 2013, 3, 879–886 RSC.
  33. L. Fan, Q. Ma, J. Tian, D. Li, X. Xi, X. Dong, W. Yu, J. Wang and G. Liu, J. Mater. Sci., 2018, 53, 2290–2302 CrossRef CAS.
  34. Z. Jiang, L. D. Tijing, A. Amarjargal, C. H. Park, K.-J. An, H. K. Shon and C. S. Kim, Composites, Part B, 2015, 77, 311–318 CrossRef CAS.
  35. L. Yang, Q. Ma, X. Xi, D. Li, J. Liu, X. Dong, W. Yu, J. Wang and G. Liu, Chem. Eng. J., 2019, 361, 713–724 CrossRef CAS.
  36. Z. Wang, Q. Ma, X. Dong, D. Li, X. Xi, W. Yu, J. Wang and G. Liu, Phys. Chem. Chem. Phys., 2017, 19, 118–126 RSC.
  37. C.-H. Park, S.-J. Kang, L. D. Tijing, H. R. Pant and C. S. Kim, Ceram. Int., 2013, 39, 9785–9790 CrossRef CAS.
  38. J. Li, X. Chen, D. Xu and K. Pan, Ecotoxicol. Environ. Saf., 2019, 170, 716–721 CrossRef CAS.
  39. D. Ponnamma, O. Aljarod, H. Parangusan and M. A. A. Al-Maadeed, Results Phys., 2019, 13, 102130 CrossRef.
  40. B. J. Tefft, S. Uthamaraj, J. J. Harburn, O. Hlinomaz, A. Lerman, D. Dragomir-Daescu and G. S. Sandhu, J. Magn. Magn. Mater., 2017, 427, 100–104 CrossRef CAS PubMed.
  41. H. Wang, H. Tang, J. He and Q. Wang, Mater. Res. Bull., 2009, 44, 1676–1680 CrossRef CAS.
  42. N. Sharma, G. H. Jaffari, S. I. Shah and D. J. Pochan, Nanotechnology, 2010, 21, 085707 CrossRef.
  43. Y. K. Sung, B. W. Ahn and T. J. Kang, J. Magn. Magn. Mater., 2012, 324, 916–922 CrossRef CAS.
  44. H. Wang, Y. Li, L. Sun, Y. Li, W. Wang, S. Wang, S. Xu and Q. Yang, J. Colloid Interface Sci., 2010, 350, 396–401 CrossRef CAS.
  45. S. Bharathkumar, M. Sakar, R. K. Vinod and S. Balakumar, Phys. Chem. Chem. Phys., 2015, 17, 17745–17754 RSC.
  46. K. Jain, S. Pathak and R. P. Pant, RSC Adv., 2016, 6, 70943–70946 RSC.
  47. G. Wang, Q. Ma, J. Tian, L. Fan, D. Li, X. Dong, W. Yu, J. Wang and G. Liu, RSC Adv., 2019, 9, 206–214 RSC.
  48. G. Sreenivasulu, J. Zhang, R. Zhang, M. Popov, V. Petrov and G. Srinivasan, Materials, 2018, 11, 18 CrossRef PubMed.
  49. S.-B. Lee, S.-J. Hong, J.-W. Yi, S.-K. Lee, Y.-H. Choa and J.-B. Kim, Met. Mater. Int., 2012, 18, 165–170 CrossRef CAS.
  50. C. Doepke, T. Grothe, P. Steblinski, M. Kloecker, L. Sabantina, D. Kosmalska, T. Blachowicz and A. Ehrmann, Nanomaterials, 2019, 9, 92 CrossRef.
  51. H. J. Wang, B. Liu, W. L. Huang, Z. Lin, J. Luo, Y. Li, L. Zhuang, W. Wang and L. L. Jiang, J. Magn. Magn. Mater., 2018, 452, 1–4 CrossRef CAS.
  52. K. P. Matabola and R. M. Moutloali, J. Mater. Sci., 2013, 48, 5475–5482 CrossRef CAS.
  53. V. Beachley and X. Wen, Mater. Sci. Eng., C, 2009, 29, 663–668 CrossRef CAS.
  54. R. Ghelich, M. K. Rad and A. A. Youzbashi, J. Eng. Fibers Fabr., 2015, 10, 12–19 Search PubMed.
  55. S. Sheng, Q. Ma, X. Dong, N. Lv, J. Wang, W. Yu and G. Liu, J. Mater. Sci.: Mater. Electron., 2014, 25, 2279–2286 CrossRef CAS.
  56. L. Fan, Q. Ma, J. Tian, D. Li, X. Xi, X. Dong, W. Yu, J. Wang and G. Liu, RSC Adv., 2017, 7, 48702–48711 RSC.
  57. C. Huang, S. J. Soenen, J. Rejman, J. Trekker, C. Liu, L. Lagae, W. Ceelen, C. Wilhelm, J. Demeester and S. C. De Smedt, Adv. Funct. Mater., 2012, 22, 2479–2486 CrossRef CAS.
  58. Y. Zhong, V. Leung, L. Y. Wan, S. Dutz, F. K. Ko and U. O. Haefeli, J. Magn. Magn. Mater., 2015, 380, 330–334 CrossRef CAS.
  59. T. I. Shalaby, M. F. El-Kady, A. E. H. M. Zaki and S. M. El-Kholy, Water Sci. Technol.: Water Supply, 2017, 17, 176–187 CAS.
  60. L. Li, F. Wang, Y. Lv, J. Liu, H. Bian, W. Wang, Y. Li and Z. Shao, ACS Omega, 2018, 3, 4220–4230 CrossRef CAS.
  61. S. Harsojo, A. F. Waluyo, H. Sosiati and K. Triyana, Adv. Mater. Res., 2015, 1123, 237–240 Search PubMed.
  62. T. Gong, W. Li, H. Chen, L. Wang, S. Shao and S. Zhou, Acta Biomater., 2012, 8, 1248–1259 CrossRef CAS PubMed.
  63. H. Shao, W. Yu, Q. Ma, X. Wang, X. Dong, Z. Liu, J. Wang, G. Liu and L. Chang, RSC Adv., 2017, 7, 32850–32860 RSC.
  64. H. Shao, Q. Ma, X. Dong, W. Yu, M. Yang, Y. Yang, J. Wang and G. Liu, Sci. Rep., 2015, 5, 14052 CrossRef CAS.
  65. J. Tian, Q. Ma, W. Yu, X. Dong, Y. Yang, B. Zhao, J. Wang and G. Liu, New J. Chem., 2017, 41, 13983–13992 RSC.
  66. H. Shao, Q. Ma, W. Yu, X. Wang, X. Dong, Z. Liu, J. Wang, G. Liu and L. Chang, J. Mater. Sci.: Mater. Electron., 2017, 28, 16762–16775 CrossRef CAS.
  67. J. Li, Q. Ma, X. Dong, D. Li, X. Xi, W. Yu, J. Wang and G. Liu, RSC Adv., 2016, 6, 96084–96092 RSC.
  68. L. Han, Q. Ma and X. Dong, RSC Adv., 2015, 5, 95674–95681 RSC.
  69. X. Zhou, Q. Ma, X. Dong, J. Wang, W. Yu and G. Liu, RSC Adv., 2015, 5, 35948–35957 RSC.
  70. X. Xi, Q. Ma, X. Dong, D. Li, W. Yu, J. Wang and G. Liu, J. Mater. Sci.: Mater. Electron., 2018, 29, 7119–7129 CrossRef CAS.
  71. B. W. Ahn and T. J. Kang, J. Appl. Polym. Sci., 2012, 125, 1567–1575 CrossRef CAS.
  72. F. Liu, Q.-Q. Ni and Y. Murakami, Text. Res. J., 2013, 83, 510–518 CrossRef.
  73. I. Savva, D. Constantinou, O. Marinica, E. Vasile, L. Vekas and T. Krasia-Christoforou, J. Magn. Magn. Mater., 2014, 352, 30–35 CrossRef CAS.
  74. Q. Ma, J. Wang, X. Dong, W. Yu and G. Liu, ChemPlusChem, 2014, 79, 290–297 CrossRef CAS.
  75. F. Bi, X. Dong, J. Wang and G. Liu, RSC Adv., 2015, 5, 12571–12577 RSC.
  76. R. Murillo-Ortiz, M. Mirabal-Garcia, J. M. Martinez-Huerta, J. G. Cabal Velarde, I. E. Castaneda-Robles and A. Lobo-Guerrero, J. Appl. Phys., 2018, 123, 105108 CrossRef.
  77. W. S. Khan, R. Asmatulu, S. Davuluri and V. K. Dandin, J. Mater. Sci. Technol., 2014, 30, 854–859 CrossRef CAS.
  78. N. A. Erfan, N. Barakat and B. J. Muller-Borer, Colloids Surf., A, 2019, 576, 63–72 CrossRef CAS.
  79. J. Y. Han, W. B. Im, D. Kim, S. H. Cheong, G. Y. Lee and D. Y. Jeon, J. Mater. Chem., 2012, 22, 5374–5381 RSC.
  80. W. Ahmad, L. J. Zhang and Y. S. Zhou, CrystEngComm, 2014, 16, 3521–3531 RSC.
  81. K. Diest, J. A. Dionne, M. Spain and H. A. Atwater, Nano Lett., 2009, 9, 2579–2583 CrossRef CAS.
  82. S. Y. Kim, K. Woo, K. Lim, K. Lee and H. S. Jang, Nanoscale, 2013, 5, 9255–9263 RSC.
  83. G. Gai, L. Wang, X. Dong, C. Zheng, W. Yu, J. Wang and X. Xiao, J. Nanopart. Res., 2013, 15, 1539 CrossRef.
  84. G. Gai, L. Wang, X. Dong and S. Xu, J. Mater. Sci., 2013, 48, 5140–5147 CrossRef CAS.
  85. K. Lai, W. Jiang, J. Z. Tang, Y. Wu, B. He, G. Wang and Z. Gu, RSC Adv., 2012, 2, 13007–13017 RSC.
  86. Q. Ma, W. Yu, X. Dong, J. Wang and G. Liu, Nanoscale, 2014, 6, 2945–2952 RSC.
  87. N. Lv, Q. Ma, X. Dong, J. Wang, W. Yu and G. Liu, ChemPlusChem, 2014, 79, 690–697 CrossRef CAS.
  88. D. Zhang, X.-Z. Jin, T. Huang, N. Zhang, X.-D. Qi, J.-H. Yang, Z.-W. Zhou and Y. Wang, ACS Appl. Mater. Interfaces, 2019, 11, 5073–5083 CrossRef CAS.
  89. J. Meng, Y. Zhang, X. Qi, H. Kong, C. Wang, Z. Xu, S. Xie, N. Gu and H. Xu, Nanoscale, 2010, 2, 2565–2569 RSC.
  90. H. M. Khanlou, B. C. Ang, S. Talebian, M. M. Barzani, M. Silakhori and H. Fauzi, Measurement, 2015, 70, 179–187 CrossRef.
  91. J. Zhu, S. Wei, X. Chen, A. B. Karki, D. Rutman, D. P. Young and Z. Guo, J. Phys. Chem. C, 2010, 114, 8844–8850 CrossRef CAS.
  92. Z. Zhou, L. Yang, J. Gao and X. Chen, Adv. Mater., 2019, 31, 1804567 CrossRef PubMed.
  93. Z. X. Jiang, Y. Zhao and P. Yang, Mater. Chem. Phys., 2018, 214, 1–7 CrossRef CAS.
  94. I. H. Chen, C.-C. Wang and C.-Y. Chen, Carbon, 2010, 48, 604–611 CrossRef CAS.
  95. Q. Wang, Y. Geng, J. Li, M. Yin, Y. Hu, Y. Liu and K. Pan, Nanotechnology, 2018, 29, 135702 CrossRef.
  96. R. Goncalves, P. Martins, X. Moya, M. Ghidini, V. Sencadas, G. Botelho, N. D. Mathur and S. Lanceros-Mendez, Nanoscale, 2015, 7, 8058–8061 RSC.
  97. D. Ghanbari, M. Salavati-Niasari, F. Beshkar and O. Amiri, J. Mater. Sci.: Mater. Electron., 2015, 26, 8358–8366 CrossRef CAS.
  98. X. Chen, S. Wei, C. Gunesoglu, J. Zhu, C. S. Southworth, L. Sun, A. B. Karki, D. P. Young and Z. Guo, Macromol. Chem. Phys., 2010, 211, 1775–1783 CrossRef CAS.
  99. P. Gupta, R. Asmatulu, R. Claus and G. Wilkes, J. Appl. Polym. Sci., 2006, 100, 4935–4942 CrossRef CAS.
  100. M. K. Shin, S. I. Kim, S. J. Kim, S. Y. Park, Y. H. Hyun, Y. Lee, K. E. Lee, S.-S. Han, D.-P. Jang, Y.-B. Kim, Z.-H. Cho, I. So and G. M. Spinks, Langmuir, 2008, 24, 12107–12111 CrossRef CAS PubMed.
  101. T. Song, Y. Z. Zhang, T. J. Zhou, C. T. Lim, S. Ramakrishna and B. Liu, Chem. Phys. Lett., 2005, 415, 317–322 CrossRef CAS.
  102. S. Hatamie, F. Mohamadyar-Toupkanlou, S. Mirzaei, M. M. Ahadian, S. Hosseinzadeh, M. Soleimani, W.-J. Sheu, Z. H. Wei, T.-F. Hsieh, W.-C. Chang and C.-L. Wang, Spin, 2019, 9, 1940011 CrossRef CAS.
  103. X. Chen, K. M. Unruh, C. Ni, B. Ali, Z. Sun, Q. Lu, J. Deitzel and J. Q. Xiao, J. Phys. Chem. C, 2011, 115, 373–378 CrossRef CAS.
  104. Y. Wang, C. Dong, Y. Chuai and Y. Wang, J. Wuhan Univ. Technol., Mater. Sci. Ed., 2015, 30, 1–5 CrossRef.
  105. Z. Zhang, Z. Li, J. Zhang, H. Bian, T. Wang, J. Gao and J. Li, Ceram. Int., 2019, 45, 457–461 CrossRef CAS.
  106. R. E. Lu, K. G. Chang, B. Fu, Y. J. Shen, M. W. Xu, S. Yang, X. P. Song, M. Liu and Y. D. Yang, J. Mater. Chem. C, 2014, 2, 8578–8584 RSC.
  107. A. Amarjargal, L. D. Tijing, C.-H. Park, I.-T. Im and C. S. Kim, Eur. Polym. J., 2013, 49, 3796–3805 CrossRef CAS.
  108. Q.-Z. Yu, Mater. Sci. Eng., B, 2010, 167, 26–30 CrossRef CAS.
  109. Y.-H. Chen, C.-H. Cheng, W.-J. Chang, Y.-C. Lin, F.-H. Lin and J.-C. Lin, Mater. Sci. Eng., C, 2016, 62, 338–349 CrossRef CAS.
  110. F. Hong, C. Yan, Y. Si, J. He, J. Yu and B. Ding, ACS Appl. Mater. Interfaces, 2015, 7, 20200–20207 CrossRef CAS.
  111. C. Han, Q. Ma, X. Dong, W. Yu, J. Wang and G. Liu, J. Mater. Sci.: Mater. Electron., 2015, 26, 2457–2465 CrossRef CAS.
  112. R. Zhao, X. Li, Y. Li, Y. Li, B. Sun, N. Zhang, S. Chao and C. Wang, J. Colloid Interface Sci., 2017, 505, 1018–1030 CrossRef CAS PubMed.
  113. L. Wu, W. Sui, C. Dong, C. Zhang and C. Jiang, Appl. Surf. Sci., 2016, 384, 368–375 CrossRef CAS.
  114. Z. Wang, X. Liu, M. Lv, P. Chai, Y. Liu, X. Zhou and J. Meng, J. Phys. Chem. C, 2008, 112, 15171–15175 CrossRef CAS.
  115. Z. Wang, X. Liu, M. Lv, P. Chai, Y. Liu and J. Meng, J. Phys. Chem. B, 2008, 112, 11292–11297 CrossRef CAS.
  116. P. Jing, J. Du, J. Wang, J. Wei, L. Pan, J. Li and Q. Liu, Sci. Rep., 2015, 5, 131 Search PubMed.
  117. C. J. Luo, E. Stride and M. Edirisinghe, Macromolecules, 2012, 45, 4669–4680 CrossRef CAS.
  118. J. Xue, T. Wu, Y. Dai and Y. Xia, Chem. Rev., 2019, 119, 5298–5415 CrossRef CAS.
  119. P. Jing, J. Du, C. Jin, J. Wang, L. Pan, J. Li and Q. Liu, J. Mater. Sci., 2016, 51, 885–892 CrossRef CAS.
  120. N. S. Hansen, D. Cho and Y. L. Joo, Small, 2012, 8, 1510–1514 CrossRef CAS PubMed.
  121. I. Abdalla, J. Shen, J. Yu, Z. Li and B. Ding, Sci. Rep., 2018, 8, 12402 CrossRef PubMed.
  122. J. Xiang, Y. Chu, X. Zhang and X. Shen, Appl. Surf. Sci., 2012, 263, 320–325 CrossRef CAS.
  123. X. L. Zhang, W. W. Pan, J. Dong, Q. F. Liu and J. B. Wang, J. Mater. Sci., 2015, 50, 7218–7226 CrossRef CAS.
  124. J. Xiang, X. Zhang, J. Li, Y. Chu and X. Shen, Chem. Phys. Lett., 2013, 576, 39–43 CrossRef CAS.
  125. D. Li, Q. Ma, Y. Song, X. Xi, X. Dong, W. Yu, J. Wang and G. Liu, Phys. Chem. Chem. Phys., 2016, 18, 27536–27544 RSC.
  126. P. P. Jing, M. T. Liu, Y. P. Pu, Y. F. Cui, Z. Wang, J. B. Wang and Q. F. Liu, Sci. Rep., 2016, 6, 37701 CrossRef CAS PubMed.
  127. Y. Ji, X. B. Zhang, Y. J. Zhu, B. Li, Y. Wang, J. C. Zhang and Y. Feng, Mater. Res. Bull., 2013, 48, 2426–2429 CrossRef CAS.
  128. X. Y. Zhang, Y. Liu, J. Li and X. J. Yang, Mater. Res. Innovations, 2013, 17, 436–439 CrossRef CAS.
  129. W. Zhang, H. P. Li and W. Pan, in High-Performance Ceramics Vii, Pts 1 and 2, ed. W. Pan and J. H. Gong, 2012, vol. 512–515, pp. 1438–1441 Search PubMed.
  130. W. Zhang, H. P. Li and W. Pan, J. Mater. Sci., 2012, 47, 8216–8222 CrossRef CAS.
  131. W. Pan, R. Han, X. Chi, Q. Liu and J. Wang, J. Alloys Compd., 2013, 577, 192–194 CrossRef CAS.
  132. Z. Chen, Y. Gu, P. Aprelev, K. Kornev, I. Luzinov, J. Chen and F. Peng, J. Am. Ceram. Soc., 2016, 99, 1504–1511 CrossRef CAS.
  133. S. H. Xie, J. Y. Li, R. Proksch, Y. M. Liu, Y. C. Zhou, Y. Y. Liu, Y. Ou, L. N. Lan and Y. Qiao, Appl. Phys. Lett., 2008, 93, 222904 CrossRef.
  134. J. Guo, Y. Yang, W. Yu, X. Dong, J. Wang, G. Liu and T. Wang, RSC Adv., 2016, 6, 111447 RSC.
  135. J. C. Fu, J. L. Zhang, C. H. Zhao, Y. Peng, X. D. Li, Y. M. He, Z. X. Zhang, X. J. Pan, N. J. Mellors and E. Q. Xie, J. Alloys Compd., 2013, 577, 97–102 CrossRef CAS.
  136. J. Xiang, X. Shen, F. Song, M. Liu, G. Zhou and Y. Chu, Mater. Res. Bull., 2011, 46, 258–261 CrossRef CAS.
  137. C. Gao, Z. Pui, C. Zhang and Q. Gao, Mater. Res. Express, 2019, 6, 025049 CrossRef.
  138. X. Q. Shen, M. Q. Liu, F. Z. Song and X. F. Meng, J. Sol-Gel Sci. Technol., 2010, 53, 448–453 CrossRef CAS.
  139. M. Sakar, S. Balakumar, P. Saravanan and S. N. Jaisankar, Mater. Des., 2016, 94, 487–495 CrossRef CAS.
  140. S. Xie, F. Ma, Y. Liu and J. Li, Nanoscale, 2011, 3, 3152–3158 RSC.
  141. X. Qu, N. Kobayashi and T. Komatsu, ACS Nano, 2010, 4, 1732–1738 CrossRef CAS PubMed.
  142. I. Abdalla, J. Yu, Z. Li and B. Ding, Composites, Part B, 2018, 155, 397–404 CrossRef CAS.
  143. P. Li, F. Lv, L. Liu, L. Ding and Y. Zhang, J. Magn. Magn. Mater., 2016, 414, 105–110 CrossRef CAS.
  144. H. Chen, J. Sun, Z. Wang, Y. Zhou, Z. Lou, B. Chen, P. Wang, Z. Guo, H. Tang, J. Ma, Y. Xia, N. Gu and F. Zhang, ACS Appl. Mater. Interfaces, 2018, 10, 44279–44289 CrossRef CAS.
  145. T.-C. Lin, F.-H. Lin and J.-C. Lin, Acta Biomater., 2012, 8, 2704–2711 CrossRef CAS PubMed.
  146. P. Wang, L. Cheng, Y. Zhang and L. Zhang, J. Alloys Compd., 2017, 716, 306–320 CrossRef CAS.
  147. M. Hussain, S. S. Batool, Z. Imran, M. Ahmad, K. Rasool, M. A. Rafiq and M. M. Hasan, Sens. Actuators, B, 2014, 192, 429–438 CrossRef CAS.
  148. Q. Yu, H. Chen, P. Chen, Q. Wang, C. Lu and C. Jia, J. Mater. Sci.: Mater. Electron., 2017, 28, 2769–2774 CrossRef CAS.
  149. S. Koombhongse, W. Liu and D. H. Reneker, J. Polym. Sci., Part B: Polym. Phys., 2001, 39, 2598–2606 CrossRef CAS.
  150. M. Bayat, H. Yang, F. K. Ko, D. Michelson and A. Mei, Polymer, 2014, 55, 936–943 CrossRef CAS.
  151. H. Ziyadi, A. Heydari and S. M. Rezayat, Ceram. Int., 2014, 40, 5913–5919 CrossRef CAS.
  152. L. Zou, Q. Wang, X. Shen, Z. Wang, M. Jing and Z. Luo, Appl. Surf. Sci., 2015, 332, 674–681 CrossRef CAS.
  153. J. Dong, Y. Zhang, X. Zhang, Q. Liu and J. Wang, Mater. Lett., 2014, 120, 9–12 CrossRef CAS.
  154. J. Lee, J. Kim, D. Kim, G. Lee, Y.-B. Oh, T.-Y. Hwang, J.-H. Lim, H.-B. Cho, J. Kim and Y.-H. Choa, ACS Appl. Mater. Interfaces, 2019, 11, 26222–26227 CrossRef CAS PubMed.
  155. L. N. Pan, F. M. Gu, D. R. Cao, P. P. Jing, J. N. Li, J. B. Wang and Q. F. Liu, Appl. Phys. A: Mater. Sci. Process., 2016, 122, 583 CrossRef.
  156. F. Bi, X. Dong, J. Wang and G. Liu, J. Mater. Sci.: Mater. Electron., 2014, 25, 4259–4267 CrossRef CAS.
  157. G. Sreenivasulu, M. Popov, R. Zhang, K. Sharma, C. Janes, A. Mukundan and G. Srinivasan, Appl. Phys. Lett., 2014, 104, 052910 CrossRef.
  158. K. N. Kim, H. R. Jung and W. J. Lee, J. Photochem. Photobiol., A, 2016, 321, 257–265 CrossRef CAS.
  159. J. X. Zhao, Y. L. Cheng, X. B. Yan, D. F. Sun, F. L. Zhu and Q. J. Xue, CrystEngComm, 2012, 14, 5879–5885 RSC.
  160. Y. Cheng, B. Zou, J. Yang, C. Wang, Y. Liu, X. Fan, L. Zhu, Y. Wang, H. Ma and X. Cao, CrystEngComm, 2011, 13, 2268–2272 RSC.
  161. C. Han, Q. Ma, Y. Yang, M. Yang, W. Yu, X. Dong, J. Wang and G. Liu, J. Mater. Sci.: Mater. Electron., 2015, 26, 8054–8064 CrossRef CAS.
  162. Q. Gao, J. Luo, X. Wang, C. Gao and M. Ge, Nanoscale Res. Lett., 2015, 10, 176 CrossRef PubMed.
  163. C. Eid, D. Luneau, V. Salles, R. Asmar, Y. Monteil, A. Khoury and A. Brioude, J. Phys. Chem. C, 2011, 115, 17643–17646 CrossRef CAS.
  164. P. Jing, J. Li, L. Pan, J. Wang, X. Sun and Q. Liu, J. Hazard. Mater., 2015, 284, 163–170 CrossRef CAS.
  165. X. Li, Z. Hou, P. A. Ma, X. Zhang, C. Li, Z. Cheng, Y. Dai, J. Lian and J. Lin, RSC Adv., 2013, 3, 8517–8526 RSC.
  166. Y. Zhou, W. Song, L. Zhang and S. Tao, J. Mater. Chem. A, 2018, 6, 12298–12307 RSC.
  167. Z. Wang, L. Zhao, P. Wang, L. Guo and J. Yu, J. Alloys Compd., 2016, 687, 541–547 CrossRef CAS.
  168. Q. Gao, J. Takizawa and M. Kimura, Polymer, 2013, 54, 120–126 CrossRef CAS.
  169. N. A. M. Barakat, B. Kim and H. Y. Kim, J. Phys. Chem. C, 2009, 113, 531–536 CrossRef CAS.
  170. L. Pan, X. Zhang, J. Wang and Q. Liu, Ceram. Int., 2017, 43, 1236–1241 CrossRef CAS.
  171. W.-Y. Lee, H. J. Yun and J.-W. Yoon, J. Alloys Compd., 2014, 583, 320–324 CrossRef CAS.
  172. B. Mordina, R. K. Tiwari, D. K. Setua and A. Sharma, RSC Adv., 2015, 5, 19091–19105 RSC.
  173. J.-G. Zhao, W.-Y. Zhang, E.-Q. Xie, X.-Y. An, J.-C. Fu, Z.-J. Liu, Z.-L. Liu, Y.-J. Zhang, Y.-F. Chen and C.-Y. Zhang, NANO, 2014, 9, 1450026 CrossRef.
  174. U. Kurtan, D. Dursun, H. Aydin, M. S. Toprak, A. Baykal and A. Bozkurt, Ceram. Int., 2016, 42, 18189–18195 CrossRef CAS.
  175. J. Bhagwan, S. Rani, V. Sivasankaran, K. L. Yadav and Y. Sharma, Appl. Surf. Sci., 2017, 426, 913–923 CrossRef CAS.
  176. Y. Lu, X. C. Yang, J. L. Zhu, F. Z. Song and X. Q. Shen, Adv. Mater. Res., 2012, 399–401, 736–740 CAS.
  177. H. Liu, Z.-G. Zhang, X.-X. Wang, G.-D. Nie, J. Zhang, S.-X. Zhang, N. Cao, S.-Y. Yan and Y.-Z. Long, J. Phys. Chem. Solids, 2018, 121, 236–246 CrossRef CAS.
  178. X. Q. Shen, J. C. Zheng, X. F. Meng and Q. R. Liang, J. Wuhan Univ. Technol., Mater. Sci. Ed., 2011, 26, 384–387 CrossRef CAS.
  179. R. Yensano, S. Pinitsoontorn, V. Amornkitbamrung and S. Maensiri, J. Supercond. Novel Magn., 2014, 27, 1553–1560 CrossRef CAS.
  180. Y. Zhang, W. W. Pan, J. Dong, Z. K. Wang, Q. F. Liu and J. B. Wang, J. Phys. D: Appl. Phys., 2013, 46, 105001 CrossRef.
  181. Q. R. Liang, X. Q. Shen, F. Z. Song and M. Q. Liu, J. Mater. Sci. Technol., 2011, 27, 996–1000 CrossRef CAS.
  182. W. Ponhan and S. Maensiri, Solid State Sci., 2009, 11, 479–484 CrossRef CAS.
  183. O. Saensuk, S. Phokha, A. Bootchanont, S. Maensiri and E. Swatsitang, Ceram. Int., 2015, 41, 8133–8141 CrossRef CAS.
  184. S. Maensiri, M. Sangmanee and A. Wiengmoon, Nanoscale Res. Lett., 2009, 4, 221–228 CrossRef CAS PubMed.
  185. P. Kidkhunthod, S. Nilmoung, S. Mahakot, S. Rodporn, S. Phumying and S. Maensiri, J. Magn. Magn. Mater., 2016, 401, 436–442 CrossRef CAS.
  186. Y.-W. Ju, J.-H. Park, H.-R. Jung, S.-J. Cho and W.-J. Lee, Compos. Sci. Technol., 2008, 68, 1704–1709 CrossRef CAS.
  187. C. Luo, X. Wang, J. Wang and K. Pan, Compos. Sci. Technol., 2016, 133, 97–103 CrossRef CAS.
  188. Y. Zhu, J. C. Zhang, J. Zhai and L. Jiang, Thin Solid Films, 2006, 510, 271–274 CrossRef CAS.
  189. N. N. Liu, P. C. Du, P. Zhou, R. G. Tanguturi, Y. J. Qi, T. J. Zhang and C. Q. Zhuang, Appl. Surf. Sci., 2020, 532, 147440 CrossRef CAS.
  190. R. Nakhowong and R. Chueachot, J. Alloys Compd., 2017, 715, 390–396 CrossRef CAS.
  191. W. Ponhan and S. Maensiri, presented in part at the 3rd IEEE International Nanoelectronics Conference, City University Hong Kong, Hong Kong, 2010.
  192. A. K. Das, M. Kar and A. Srinivasan, Phys. B, 2014, 448, 112–114 CrossRef.
  193. W. Ponhan, E. Swatsitang and S. Maensiri, Mater. Sci. Technol., 2010, 26, 1298–1303 CrossRef CAS.
  194. Y. W. Ju, J. H. Park, H. R. Jung, S. J. Cho and W. J. Lee, Mater. Sci. Eng., B, 2008, 147, 7–12 CrossRef CAS.
  195. J. Li, Y. Wu, M. Yang, Y. Yuan, W. Yin, Q. Peng, Y. Li and X. He, J. Am. Ceram. Soc., 2017, 100, 5460–5470 CrossRef CAS.
  196. T. Zhang, D. Huang, Y. Yang, F. Kang and J. Gu, Mater. Sci. Eng., B, 2013, 178, 1–9 CrossRef CAS.
  197. Y. Yang, Z. Guo, H. Zhang, D. Huang, J. Gu, Z. Huang, F. Kang, T. A. Hatton and G. C. Rutledge, J. Appl. Polym. Sci., 2013, 127, 4288–4295 CrossRef CAS.
  198. M. Bayat, H. Yang and F. Ko, Polymer, 2011, 52, 1645–1653 CrossRef CAS.
  199. C. Lu, Z. Bao, C. Qin, L. Dai and A. Zhu, RSC Adv., 2016, 6, 110155 RSC.
  200. C.-J. Li, J.-N. Wang, B. Wang, J. R. Gong and Z. Lin, Mater. Res. Bull., 2012, 47, 333–337 CrossRef CAS.
  201. P. Jing, L. Pan, J. Du, J. Wang and Q. Liu, Phys. Chem. Chem. Phys., 2015, 17, 12841–12848 RSC.
  202. B. Y. Ren, W. Shen, L. Li, S. Z. Wu and W. Wang, Appl. Surf. Sci., 2018, 447, 711–723 CrossRef CAS.
  203. G. George and S. Anandhan, Mater. Sci. Semicond. Process., 2015, 32, 40–48 CrossRef CAS.
  204. B. Tarus, N. Fadel, A. Al-Oufy and M. El-Messiry, Alexandria Eng. J., 2016, 55, 2975–2984 CrossRef.
  205. V. Pillay, C. Dott, Y. E. Choonara, C. Tyagi, L. Tomar, P. Kumar, L. C. du Toit and V. M. K. Ndesendo, J. Nanomater., 2013, 2013, 789289 Search PubMed.
  206. J. Lee, T.-Y. Hwang, M. K. Kang, H.-B. Cho, J. Kim, N. V. Myung and Y.-H. Choa, Front. Chem., 2018, 6, 18 CrossRef.
  207. Q. Yang, Q. Xu, M. Sobhan, Q. Ke, F. Anariba, K. P. Ong and P. Wu, Phys. Status Solidi RRL, 2014, 8, 653–657 CrossRef CAS.
  208. C.-J. Li, B. Wang and J.-N. Wang, J. Magn. Magn. Mater., 2012, 324, 1305–1311 CrossRef CAS.
  209. M. Liu, X. Shen, F. Song, J. Xiang and X. Meng, J. Solid State Chem., 2011, 184, 871–876 CrossRef CAS.
  210. O. Saensuk, S. Maensiri, A. Bootchanont and E. Swatsitang, Microelectron. Eng., 2014, 126, 158–164 CrossRef CAS.
  211. J. Xiang, X. Shen, F. Song and M. Liu, J. Solid State Chem., 2010, 183, 1239–1244 CrossRef CAS.
  212. C.-J. Li, B.-N. Huang and J.-N. Wang, J. Mater. Sci., 2013, 48, 1702–1710 CrossRef CAS.
  213. M. Mansour, M. Bechelany, R. Habchi and C. Eid, Phys. Lett. A, 2017, 381, 658–662 CrossRef CAS.
  214. M. S. Cortes, A. Martinez-Luevanos, L. A. Garcia-Cerda, O. S. Rodriguez-Fernandez, A. F. Fuentes, J. Romero-Garcia and S. M. Montemayor, J. Alloys Compd., 2015, 653, 290–297 CrossRef CAS.
  215. J. Xiang, Y. Q. Chu, X. Q. Shen, G. Z. Zhou and Y. T. Guo, J. Colloid Interface Sci., 2012, 376, 57–61 CrossRef CAS.
  216. W. Pan, F. Gu, K. Qi, Q. Liu and J. Wang, Mater. Chem. Phys., 2012, 134, 1097–1101 CrossRef CAS.
  217. G. D. Nie, X. F. Lu, J. Y. Lei, Z. Q. Jiang and C. Wang, J. Mater. Chem. A, 2014, 2, 15495–15501 RSC.
  218. J. M. Li, X. L. Zeng, G. Q. Wu and Z. A. Xu, CrystEngComm, 2012, 14, 525–532 RSC.
  219. J. X. Sui, X. X. Wang, C. Song, Q. Liu, F. Yuan and Y. Z. Long, J. Magn. Magn. Mater., 2018, 467, 74–81 CrossRef CAS.
  220. S. Sonsupap, P. Kidkhunthod, N. Chanlek, S. Pinitsoontorn and S. Maensiri, Appl. Surf. Sci., 2016, 380, 16–22 CrossRef CAS.
  221. P. Mohanapriya, C. I. Sathish, R. Pradeepkumar, H. Segawa, K. Yamaura, K. Watanabe, T. S. Natarajan and N. V. Jaya, J. Nanosci. Nanotechnol., 2013, 13, 5391–5400 CrossRef CAS.
  222. M. C. Maldonado-Orozco, M. T. Ochoa-Lara, J. E. Sosa-Marquez, R. P. Talamantes-Soto, A. Hurtado-Macias, R. Lopez Anton, J. A. Gonzalez, J. T. Holguin-Momaca, S. F. Olive-Mendez and F. Espinosa-Magana, J. Am. Ceram. Soc., 2019, 102, 2800–2809 CrossRef CAS.
  223. S. Liu, J. Jia, J. Wang, S. Liu, X. Wang, H. Song and X. Hu, J. Magn. Magn. Mater., 2012, 324, 2070–2074 CrossRef CAS.
  224. C. Eid, E. Assaf, R. Habchi, P. Miele and M. Bechelany, RSC Adv., 2015, 5, 97849–97854 RSC.
  225. N. A. M. Barakat, M. F. Abadir, K. T. Nam, A. M. Hamza, S. S. Al-Deyab, W.-I. Baek and H. Y. Kim, J. Mater. Chem., 2011, 21, 10957–10964 RSC.
  226. F. Wang, Y. Sun, D. Li, B. Zhong, Z. Wu, S. Zuo, D. Yan, R. Zhuo, J. Feng and P. Yan, Carbon, 2018, 134, 264–273 CrossRef CAS.
  227. J. Zhang, P. Wang, Y. Chen, G. Wang, D. Wang, L. Qiao, T. Wang and F. Li, J. Electron. Mater., 2018, 47, 4703–4709 CrossRef.
  228. Y. Jiang, X. Fu, Z. Zhang, W. Du, P. Xie, C. Cheng and R. Fan, J. Alloys Compd., 2019, 804, 305–313 CrossRef CAS.
  229. I. Abdalla, A. Salim, M. Zhu, J. Yu, Z. Li and B. Ding, ACS Appl. Mater. Interfaces, 2018, 10, 44561–44569 CrossRef CAS.
  230. A. Chaudhary, S. Kumari, R. Kumar, S. Teotia, B. P. Singh, A. P. Singh, S. K. Dhawan and S. R. Dhakate, ACS Appl. Mater. Interfaces, 2016, 8, 10600–10608 CrossRef CAS.
  231. F. M. Oliveira and R. Gusmão, ACS Appl. Electron. Mater., 2020, 2, 3048–3071 CrossRef CAS.
  232. T. Wang, H. Wang, X. Chi, R. Li and J. Wang, Carbon, 2014, 74, 312–318 CrossRef CAS.
  233. F. Shahzad, M. Alhabeb, C. B. Hatter, B. Anasori, S. Man Hong, C. M. Koo and Y. Gogotsi, Science, 2016, 353, 1137–1140 CrossRef CAS.
  234. M. An, J. Cui and L. Wang, J. Phys. Chem. C, 2014, 118, 3062–3068 CrossRef CAS.
  235. W. Song, Z. Z. Yang, F. Q. Ma, M. Q. Chi, B. Zhao and X. F. Lu, RSC Adv., 2017, 7, 40334–40341 RSC.
  236. H. Zhao, L. Zhang, X. Gu, S. Li, B. Li, H. Wang, J. Yang and J. Liu, RSC Adv., 2015, 5, 10951–10959 RSC.
  237. M.-J. Chang, H. Wang, H.-L. Li, J. Liu and H.-L. Du, J. Mater. Sci., 2018, 53, 3682–3691 CrossRef CAS.
  238. A. M. El-Rafei, A. S. El-Kalliny and T. A. Gad-Allah, J. Magn. Magn. Mater., 2017, 428, 92–98 CrossRef CAS.
  239. W. Wang, N. Li, Y. Chi, Y. Li, W. Yan, X. Li and C. Shao, Ceram. Int., 2013, 39, 3511–3518 CrossRef CAS.
  240. Y.-N. Feng, H.-C. Wang, Y.-D. Luo, Y. Shen and Y.-H. Lin, J. Appl. Phys., 2013, 113, 146101 CrossRef.
  241. R. Yang, H. Sun, J. Li and Y. A. Li, Ceram. Int., 2018, 44, 14032–14035 CrossRef CAS.
  242. V. Pongsorrarith, C. Srisitthiratkul, K. Laohhasurayotin and N. Intasanta, Mater. Lett., 2012, 67, 1–4 CrossRef CAS.
  243. K. Y. A. Lin, M. T. Yang, J. T. Lin and Y. C. Du, Chemosphere, 2018, 208, 502–511 CrossRef CAS PubMed.
  244. Z. Zhang, Y. Jiang, M. Chi, Z. Yang, G. Nie, X. Lu and C. Wang, Appl. Surf. Sci., 2016, 363, 578–585 CrossRef CAS.
  245. E. Ghasemi, H. Ziyadi, A. M. Afshar and M. Sillanpaa, Chem. Eng. J., 2015, 264, 146–151 CrossRef CAS.
  246. D. Chen, Y. Li, M. Bao, Y. Hou, J. Jin, Z. Yin and Z. Wang, ACS Sustainable Chem. Eng., 2019, 7, 13786–13795 CrossRef CAS.
  247. J. J. Du and C. Y. Jing, J. Phys. Chem. C, 2011, 115, 17829–17835 CrossRef CAS.
  248. F.-C. Liang, Y.-L. Luo, C.-C. Kuo, B.-Y. Chen, C.-J. Cho, F.-J. Lin, Y.-Y. Yu and R. Borsali, Polymers, 2017, 9, 136 CrossRef PubMed.
  249. S. Zhan, D. Zhu, G. Ren, Z. Shen, M. Qiu, S. Yang, H. Yu and Y. Li, ACS Appl. Mater. Interfaces, 2014, 6, 16841–16850 CrossRef CAS PubMed.
  250. S. Al-Meer, Z. K. Ghouri, K. Elsaid, A. Easa, M. T. Al-Qahtani and M. S. Akhtar, J. Alloys Compd., 2017, 723, 477–483 CrossRef CAS.
  251. M. Venkatesan, L. Veeramuthu, F.-C. Liang, W.-C. Chen, C.-J. Cho, C.-W. Chen, J.-Y. Chen, Y. Yan, S.-H. Chang and C.-C. Kuo, Chem. Eng. J., 2020, 397, 125431 CrossRef CAS.
  252. P. Zhou, F. Liang, Y. Liu, R. Deng, H. Yang, Q. Wang, X. Qu and Z. Yang, Sci. China Mater., 2015, 58, 961–968 CrossRef CAS.
  253. N. H. A. Ngadiman, A. Idris, M. Irfan, D. Kurniawan, N. M. Yusof and R. Nasiri, J. Mech. Behav. Biomed. Mater., 2015, 49, 90–104 CrossRef CAS.
  254. J. Meng, B. Xiao, Y. Zhang, J. Liu, H. Xue, J. Lei, H. Kong, Y. Huang, Z. Jin, N. Gu and H. Xu, Sci. Rep., 2013, 3, 2655 CrossRef PubMed.
  255. N. H. A. Ngadiman, N. Mohd Yusof, A. Idris, D. Kurniawan and E. Fallahiarezoudar, J. Bioact. Compat. Polym., 2017, 32, 411–428 CrossRef CAS.
  256. R. K. Singh, K. D. Patel, J. H. Lee, E.-J. Lee, J.-H. Kim, T.-H. Kim and H.-W. Kim, PLoS One, 2014, 9, e91584 CrossRef.
  257. Q. Cai, Y. Shi, D. Shan, W. Jia, S. Duan, X. Deng and X. Yang, Mater. Sci. Eng., C, 2015, 55, 166–173 CrossRef CAS.
  258. J. T. Kannarkat, J. Battogtokh, J. Philip, O. C. Wilson and P. M. Mehl, J. Appl. Phys., 2010, 107, 09B307 CrossRef.
  259. A. Jedlovszky-Hajdu, K. Molnar, P. M. Nagy, K. Sinko and M. Zrinyi, Colloids Surf., A, 2016, 503, 79–87 CrossRef CAS.
  260. N. Shankhwar, M. Kumar, B. B. Mandal and A. Srinivasan, Mater. Sci. Eng., C, 2016, 69, 1167–1174 CrossRef CAS.
  261. J. Jose, R. Kumar, S. Harilal, G. E. Mathew, D. G. T. Parambi, A. Prabhu, M. S. Uddin, L. Aleya, H. Kim and B. Mathew, Environ. Sci. Pollut. Res., 2020, 27, 19214–19225 CrossRef.
  262. H. P. Kok, P. Wust, P. R. Stauffer, F. Bardati, G. C. van Rhoon and J. Crezee, Radiat. Oncol., 2015, 10, 196 CrossRef CAS.
  263. X. Feng, J. Li, X. Zhang, T. Liu, J. Ding and X. Chen, J. Controlled Release, 2019, 302, 19–41 CrossRef CAS.
  264. S. Thakkar and M. Misra, Eur. J. Pharm. Sci., 2017, 107, 148–167 CrossRef CAS PubMed.
  265. M. Jin, D.-G. Yu, C. F. G. C. Geraldes, G. R. Williams and S. W. A. Bligh, Mol. Pharmaceutics, 2016, 13, 2457–2465 CrossRef CAS PubMed.
  266. A. R. K. Sasikala, A. R. Unnithan, Y.-H. Yun, C. H. Park and C. S. Kim, Acta Biomater., 2016, 31, 122–133 CrossRef CAS PubMed.
  267. H. Zhang, J. Xia, X. Pang, M. Zhao, B. Wang, L. Yang, H. Wan, J. Wu and S. Fu, Mater. Sci. Eng., C, 2017, 73, 537–543 CrossRef CAS.
  268. H. J. Haroosh, Y. Dong and G. D. Ingram, J. Polym. Sci., Part B: Polym. Phys., 2013, 51, 1607–1617 CrossRef CAS.
  269. M. R. Hasan, S. H. Baek, K. S. Seong, J. H. Kim and I. K. Park, ACS Appl. Mater. Interfaces, 2015, 7, 5768–5774 CrossRef CAS PubMed.
  270. F.-R. Fan, Z.-Q. Tian and Z. Lin Wang, Nano Energy, 2012, 1, 328–334 CrossRef CAS.
  271. P. Glynne-Jones, S. Beeby and N. White, IEE Proc.-A: Sci., Meas. Technol., 2001, 148, 68–72 Search PubMed.
  272. X. Chen, Y. Wu, A. Yu, L. Xu, L. Zheng, Y. Liu, H. Li and Z. Lin Wang, Nano Energy, 2017, 38, 91–100 CrossRef CAS.
  273. K. Dong, J. Deng, Y. Zi, Y.-C. Wang, C. Xu, H. Zou, W. Ding, Y. Dai, B. Gu, B. Sun and Z. L. Wang, Adv. Mater., 2017, 29, 1702648 CrossRef.
  274. G. Zhu, C. Pan, W. Guo, C.-Y. Chen, Y. Zhou, R. Yu and Z. L. Wang, Nano Lett., 2012, 12, 4960–4965 CrossRef CAS.
  275. Z. L. Wang, ACS Nano, 2013, 7, 9533–9557 CrossRef CAS.
  276. W. Zeng, L. Shu, Q. Li, S. Chen, F. Wang and X.-M. Tao, Adv. Mater., 2014, 26, 5310–5336 CrossRef CAS PubMed.
  277. Y. Mao, P. Zhao, G. McConohy, H. Yang, Y. Tong and X. Wang, Adv. Energy Mater., 2014, 4, 1301624 CrossRef.
  278. Z. Pi, J. Zhang, C. Wen, Z.-B. Zhang and D. Wu, Nano Energy, 2014, 7, 33–41 CrossRef CAS.
  279. S. Cheon, H. Kang, H. Kim, Y. Son, J. Y. Lee, H.-J. Shin, S.-W. Kim and J. Cho, Adv. Funct. Mater., 2017, 28, 1703778 CrossRef.
  280. R. A. Whiter, V. Narayan and S. Kar-Narayan, Adv. Energy Mater., 2014, 4, 1400519 CrossRef.
  281. D. Ponnamma, O. Aljarod, H. Parangusan and M. A. Al-Maadeed, Mater. Chem. Phys., 2020, 239, 122257 CrossRef CAS.
  282. J.-S. Im and I.-K. Park, ACS Appl. Mater. Interfaces, 2018, 10, 25660–25665 CrossRef CAS PubMed.
  283. W. Wu, S. Bai, M. Yuan, Y. Qin, Z. L. Wang and T. Jing, ACS Nano, 2012, 6, 6231–6235 CrossRef CAS PubMed.
  284. A. Ehrmann and T. Blachowicz, J. Magn. Magn. Mater., 2019, 475, 727 CrossRef CAS.
  285. N. Kikuchi, S. Okamoto, O. Kitakami, Y. Shimada, S. G. Kim, Y. Otani and K. Fukamichi, J. Appl. Phys., 2001, 90, 6548–6549 CrossRef CAS.
  286. B. Van Waeyenberge, A. Puzic, H. Stoll, K. W. Chou, T. Tyliszczak, R. Hertel, M. Faehnle, H. Brueckl, K. Rott, G. Reiss, I. Neudecker, D. Weiss, C. H. Back and G. Schuetz, Nature, 2006, 444, 461–464 CrossRef CAS PubMed.
  287. M. Weigand, B. Van Waeyenberge, A. Vansteenkiste, M. Curcic, V. Sackmann, H. Stoll, T. Tyliszczak, K. Kaznatcheev, D. Bertwistle, G. Woltersdorf, C. H. Back and G. Schuetz, Phys. Rev. Lett., 2009, 102, 077201 CrossRef PubMed.
  288. T. Blachowicz, C. Dopke and A. Ehrmann, Nanomaterials, 2020, 10, 738 CrossRef CAS PubMed.
  289. H.-R. Kim, B.-S. Kim and I.-S. Kim, Mater. Chem. Phys., 2012, 135, 1024–1029 CrossRef CAS.
  290. E. Fallahiarezoudar, M. Ahmadipourroudposht, N. M. Yusof and A. Idris, Int. J. Polym. Mater. Polym. Biomater., 2015, 64, 15–24 CrossRef CAS.
  291. A. F. Lubambo, L. Ono, V. Drago, N. Mattoso, J. Varalda, M. R. Sierakowski, C. N. Sakakibara, R. A. Freitas and C. K. Saul, Carbohydr. Polym., 2015, 134, 775–783 CrossRef CAS PubMed.
  292. L. Hou, W. M. R. N. Udansawa, A. Pochiraju, W. Dong, Y. Zheng, R. J. Linhardt and T. J. Simmons, ACS Biomater. Sci. Eng., 2016, 2, 1905–1913 CrossRef CAS PubMed.
  293. L. Li, F. Wang, Y. Lv, J. Liu, D. Zhang and Z. Shao, Appl. Clay Sci., 2018, 161, 225–234 CrossRef CAS.
  294. T.-C. Lin, F.-H. Lin and J.-C. Lin, J. Biomater. Sci., Polym. Ed., 2013, 24, 1152–1163 CrossRef CAS PubMed.
  295. J. G. Duran-Guerrero, M. A. Martinez-Rodriguez, M. A. Garza-Navarro, V. A. Gonzalez-Gonzalez, A. Torres-Castro and J. R. De La Rosa, Carbohydr. Polym., 2018, 200, 289–296 CrossRef CAS PubMed.
  296. Y. Wei, X. Zhang, Y. Song, B. Han, X. Hu, X. Wang, Y. Lin and X. Deng, Biomed. Mater., 2011, 6, 055008 CrossRef PubMed.
  297. T. Huang, P. Song, L. Jiang, Y. Peng, S. Feng and J. Wang, Electrophoresis, 2016, 37, 2050–2053 CrossRef CAS PubMed.
  298. D. Navarathne, Y. Ner, M. Jain, J. G. Grote and G. A. Sotzing, Mater. Lett., 2011, 65, 219–221 CrossRef CAS.
  299. A. GhavamiNejad, A. R. K. Sasikala, A. R. Unnithan, R. G. Thomas, Y. Y. Jeong, M. Vatankhah-Varnoosfaderani, F. J. Stadler, C. H. Park and C. S. Kim, Adv. Funct. Mater., 2015, 25, 2867–2875 CrossRef CAS.
  300. J. Guo, X. Ye, W. Liu, Q. Wu, H. Shen and K. Shu, Mater. Lett., 2009, 63, 1326–1328 CrossRef CAS.
  301. M. S. A. Darwish, A. Bakry, L. M. Al-Harbi, M. M. Khowdiary, A. A. El-Henawy and J. Yoon, J. Dispersion Sci. Technol., 2020, 41, 1711–1719 CrossRef CAS.
  302. M. S. A. Darwish, A. Bakry, O. Kolek, L. Martinova and I. Stibor, Polym. Compos., 2019, 40, 296–303 CrossRef CAS.
  303. R. Mincheva, O. Stoilova, H. Penchev, T. Ruskov, I. Spirov, N. Manolova and I. Rashkov, Eur. Polym. J., 2008, 44, 615–627 CrossRef CAS.
  304. D. Zhang, A. B. Karki, D. Rutman, D. R. Young, A. Wang, D. Cocke, T. H. Ho and Z. Guo, Polymer, 2009, 50, 4189–4198 CrossRef CAS.
  305. B. Wang, Y. Sun and H. Wang, J. Appl. Polym. Sci., 2010, 115, 1781–1786 CrossRef CAS.
  306. A. Sindelo and T. Nyokong, Heliyon, 2019, 5, e02352 CrossRef PubMed.
  307. Z. Wang, Q. Ma, X. Dong, D. Li, X. Xi, W. Yu, J. Wang and G. Liu, ACS Appl. Mater. Interfaces, 2016, 8, 26226–26234 CrossRef CAS PubMed.
  308. S. Cavaliere, V. Salles, A. Brioude, Y. Lalatonne, L. Motte, P. Monod, D. Cornu and P. Miele, J. Nanopart. Res., 2010, 12, 2735–2740 CrossRef CAS.
  309. D. Li, Y. Gu, X. Xu, Y. Feng, Y. Ma, S. Li and C. Yao, Microchim. Acta, 2019, 186, 542 CrossRef CAS PubMed.
  310. X.-Y. Ye, Z.-M. Liu, Z.-G. Wang, X.-J. Huang and Z.-K. Xu, Mater. Lett., 2009, 63, 1810–1813 CrossRef CAS.
  311. N. Sabzroo, T. R. Bastami, M. Karimi, T. Heidari, S. Agarwal and V. K. Gupta, J. Ind. Eng. Chem., 2018, 60, 237–249 CrossRef CAS.
  312. H. Bagheri, M. N. Mobara, A. Roostaie and M. Y. Baktash, J. Sep. Sci., 2017, 40, 3857–3865 CrossRef CAS.
  313. H. Bagheri, P. Khanipour and S. Asgari, Anal. Chim. Acta, 2016, 934, 88–97 CrossRef CAS PubMed.
  314. P. Tiwari, S. Agarwal, S. Srivastava and S. Jain, J. Biomed. Mater. Res., Part B, 2018, 106, 40–51 CrossRef CAS.
  315. W. Xue, Y. Hu, F. Wang, X. Yang and L. Wang, Colloids Surf., A, 2019, 564, 115–121 CrossRef CAS.
  316. D. Demir, D. Gures, T. Tecim, R. Genc and N. Bolgen, Appl. Nanosci., 2018, 8, 1461–1469 CrossRef CAS.
  317. B. Wang, H. Zheng, M.-W. Chang, Z. Ahmad and J.-S. Li, Colloids Surf., B, 2016, 145, 757–767 CrossRef CAS PubMed.
  318. S. Wu, B. Wang, Z. Ahmad, J. Huang, M.-W. Chang and J.-S. Li, Mater. Lett., 2017, 204, 73–76 CrossRef CAS.
  319. M. Hadjian, D. Semnani and J. Varshosaz, Polym. Adv. Technol., 2019, 30, 2729–2741 CrossRef.
  320. Q. Yu, M. Ma, P. Chen, Q. Wang, C. Lu, Y. Gao, R. Wang and H. Chen, Polym. Eng. Sci., 2017, 57, 1186–1192 CAS.
  321. A. Wang, H. Singh, T. A. Hatton and G. C. Rutledge, Polymer, 2004, 45, 5505–5514 CrossRef.
  322. I. Savva, O. Marinica, C. A. Papatryfonos, L. Vekas and T. Krasia-Christoforou, RSC Adv., 2015, 5, 16484–16496 RSC.
  323. L. Burke, C. J. Mortimer, D. J. Curtis, A. R. Lewis, R. Williams, K. Hawkins, T. G. G. Maffeis and C. J. Wright, Mater. Sci. Eng., C, 2017, 70, 512–519 CrossRef CAS PubMed.
  324. D. Guo, Z. Sun, L. Xu, Y. Gao, M. Dai, S. Wang, Q. Chang, C. Wang and D. Ma, Mater. Lett., 2015, 159, 159–162 CrossRef CAS.
  325. E. Korina, O. Stoilova, N. Manolova and I. Rashkov, J. Mater. Sci., 2014, 49, 2144–2153 CrossRef CAS.
  326. M.-H. Ho, S.-Y. Li, C.-Y. Ciou and T.-M. Wu, J. Appl. Polym. Sci., 2014, 131, 41070 Search PubMed.
  327. S. T. Tan, J. H. Wendorff, C. Pietzonka, Z. H. Jia and G. Q. Wang, ChemPhysChem, 2005, 6, 1461–1465 CrossRef CAS PubMed.
  328. M. Kumar, D. Unruh, R. Sindelar and F. Renz, Nano Hybrids, 2017, 14, 39–47 Search PubMed.
  329. D. Shan, Y. Shi, S. Duan, Y. Wei, Q. Cai and X. Yang, Mater. Sci. Eng., C, 2013, 33, 3498–3505 CrossRef CAS PubMed.
  330. X. Xi, W. Yu, Q. Ma, D. Li, X. Dong, J. Wang and G. Liu, RSC Adv., 2018, 8, 31608–31617 RSC.
  331. H. Shao, Q. Ma, X. Dong, W. Yu, M. Yang, Y. Yang, J. Wang and G. Liu, Phys. Chem. Chem. Phys., 2015, 17, 21845–21855 RSC.
  332. J. Tian, Q. Ma, W. Yu, D. Li, X. Dong, G. Liu and J. Wang, RSC Adv., 2019, 9, 10679–10692 RSC.
  333. Y. Xie, Q. Ma, H. Qi, Y. Song, J. Tian, M. A. Musa, W. Yu, X. Dong, D. Li and G. Liu, RSC Adv., 2019, 9, 30890–30904 RSC.
  334. Y. Xie, Q. Ma, H. Qi, Y. Song, J. Tian, W. Yu, X. Dong, D. Li, G. Liu and J. Wang, Nanotechnology, 2019, 30, 435602 CrossRef CAS.
  335. X. Xi, W. Yu, D. Li, Q. Ma, X. Dong, J. Wang and G. Liu, New J. Chem., 2018, 42, 18708–18716 RSC.
  336. Q. Ma, J. Wang, X. Dong, W. Yu and G. Liu, Chem. Eng. J., 2015, 260, 222–230 CrossRef CAS.
  337. J. Tian, Q. Ma, W. Yu, D. Li, X. Dong, G. Liu and J. Wang, Mater. Des., 2019, 170, 107701 CrossRef CAS.
  338. D. Yin, Q. Ma, X. Dong, N. Lv, J. Wang, W. Yu and G. Liu, ChemPlusChem, 2015, 80, 568–575 CrossRef CAS PubMed.
  339. S.-R. Huang, K.-F. Lin, T.-M. Don, W.-Y. Chiu and M.-F. Lin, J. Polym. Sci., Part A: Polym. Chem., 2015, 53, 2152–2162 CrossRef CAS.
  340. A. Feizbakiish and S. Ehteshami, J. AOAC Int., 2017, 100, 198–205 CrossRef.
  341. B. Song, J. Zhu and H. Fan, Mar. Pollut. Bull., 2017, 120, 159–164 CrossRef CAS.
  342. S. Wang, C. Wang, B. Zhang, Z. Sun, Z. Li, X. Jiang and X. Bai, Mater. Lett., 2010, 64, 9–11 CrossRef CAS.
  343. X. Huang, R. Wang, T. Jiao, G. Zou, F. Zhan, J. Yin, L. Zhang, J. Zhou and Q. Peng, ACS Omega, 2019, 4, 1897–1906 CrossRef CAS.
  344. E. Korina, O. Stoilova, N. Manolova and I. Rashkov, J. Environ. Chem. Eng., 2018, 6, 2075–2084 CrossRef CAS.
  345. O. Chiscan, I. Dumitru, P. Postolache, V. Tura and A. Stancu, Mater. Lett., 2012, 68, 251–254 CrossRef CAS.
  346. T. Zheng, Z. Yue, G. G. Wallace, Y. Du, P. Martins, S. Lanceros-Mendez and M. J. Higgins, Nanotechnology, 2017, 28, 065707 CrossRef PubMed.
  347. V. Kalra, J. Lee, J. H. Lee, S. G. Lee, M. Marquez, U. Wiesner and Y. L. Joo, Small, 2008, 4, 2067–2073 CrossRef CAS PubMed.
  348. Y. Liu, Q. Ma, X. Dong, W. Yu, J. Wang and G. Liu, Phys. Chem. Chem. Phys., 2015, 17, 22977–22984 RSC.
  349. Y. Liu, Q. Ma, M. Yang, X. Dong, Y. Yang, J. Wang, W. Yu and G. Liu, Chem. Eng. J., 2016, 284, 831–840 CrossRef CAS.
  350. B. Wang, P.-p. Zhang, G. R. Williams, C. Branford-White, J. Quan, H.-l. Nie and L.-m. Zhu, J. Mater. Sci., 2013, 48, 3991–3998 CrossRef CAS.
  351. X. Xi, Q. Ma, X. Dong, J. Wang, W. Yu and G. Liu, IEEE Trans. Nanotechnol., 2015, 14, 243–249 CAS.
  352. D. Yin, Q. Ma, X. Dong, N. Lv, J. Wang, W. Yu and G. Liu, J. Mater. Sci.: Mater. Electron., 2015, 26, 2658–2667 CrossRef CAS.
  353. X. Xi, Q. Ma, X. Dong, D. Li, W. Yu, J. Wang and G. Liu, J. Mater. Sci.: Mater. Electron., 2018, 29, 10284–10300 CrossRef CAS.
  354. H. Chen, Q. Ma, J. Tian, X. Li, D. Li, X. Dong, W. Yu, J. Wang and G. Liu, New J. Chem., 2019, 43, 7984–7996 RSC.
  355. K. Nasouri and A. M. Shoushtari, J. Thermoplast. Compos. Mater., 2018, 31, 431–446 CrossRef CAS.
  356. C.-R. Lin, T.-C. Tsai, M. Chung and S.-Z. Lu, J. Appl. Phys., 2009, 105, 07B509 CrossRef.
  357. S. Sheng, Q. Ma, X. Dong, N. Lv, J. Wang, W. Yu and G. Liu, J. Mater. Sci.: Mater. Electron., 2014, 25, 1309–1316 CrossRef CAS.
  358. J. Choi, J. Kim, H. Yang, F. K. Ko and K. H. Kim, J. Magn., 2015, 20, 215–220 CrossRef.
  359. M. Chen, H. Qu, J. Zhu, Z. Luo, A. Khasanov, A. S. Kucknoor, N. Haldolaarachchige, D. R. Young, S. Wei and Z. Guo, Polymer, 2012, 53, 4501–4511 CrossRef CAS.
  360. J. Dankova, M. Buzgo, J. Vejpravova, S. Kubickova, V. Sovkova, L. Vyslouzilova, A. Mantlikova, A. Necas and E. Amler, Int. J. Nanomed., 2015, 10, 7307–7317 CrossRef CAS PubMed.
  361. L. Li, G. Yang, J. Li, S. Ding and S. Zhou, Mater. Sci. Eng., C, 2014, 34, 252–261 CrossRef CAS.
  362. N. Hernandez, V. A. Gonzalez-Gonzalez, I. B. Dzul-Bautista, N. Ornelas-Soto, J. M. Barandiaran and J. Gutierrez, Eur. Polym. J., 2018, 109, 336–340 CrossRef CAS.
  363. T. Song, Y. Z. Zhang and T. J. Zhou, J. Magn. Magn. Mater., 2006, 303, E286–E289 CrossRef CAS.
  364. Z. Thompson, S. Rahman, S. Yarmolenko, J. Sankar, D. Kumar and N. Bhattarai, Materials, 2017, 10, 937 CrossRef PubMed.
  365. M. Asadi, S. Shahabuddin, A. Mollahosseini, J. Kaur and R. Saidur, J. Inorg. Organomet. Polym. Mater., 2019, 29, 1057–1066 CrossRef CAS.
  366. B.-R. Koo, I.-K. Park and H.-J. Ahn, J. Alloys Compd., 2014, 603, 52–56 CrossRef CAS.
  367. C.-J. Li, B. Wang and J.-N. Wang, J. Nanosci. Nanotechnol., 2012, 12, 2522–2528 CrossRef CAS PubMed.
  368. C.-J. Li and J.-N. Wang, Mater. Lett., 2010, 64, 586–588 CrossRef CAS.
  369. G.-F. Liu, Z.-D. Zhang, F. Dang, C.-B. Cheng, C.-X. Hou and S.-D. Liu, J. Magn. Magn. Mater., 2016, 412, 55–62 CrossRef CAS.
  370. F. Bi, X. Dong, J. Wang and G. Liu, J. Mater. Sci., 2015, 50, 1679–1687 CrossRef CAS.
  371. X. Xi, Q. Ma, M. Yang, X. Dong, J. Wang, W. Yu and G. Liu, J. Mater. Sci.: Mater. Electron., 2014, 25, 4024–4032 CrossRef CAS.
  372. X. Xu, Q. Wang, H. C. Choi and Y. H. Kim, J. Membr. Sci., 2010, 348, 231–237 CrossRef CAS.
  373. T.-I. Yang and S.-H. Chang, Nanotechnology, 2017, 28, 055601 CrossRef PubMed.
  374. V. Guarino, G. Ausanio, V. Iannotti, L. Ambrosio and L. Lanotte, eXPRESS Polym. Lett., 2018, 12, 318–329 CrossRef CAS.
  375. R. Brito-Pereira, D. M. Correia, C. Ribeiro, A. Francesko, I. Etxebarria, L. Perez-Alvarez, J. L. Vilas, P. Martins and S. Lanceros-Mendez, Composites, Part B, 2018, 141, 70–75 CrossRef CAS.
  376. I. Munaweera, A. Aliev and K. J. Balkus Jr, ACS Appl. Mater. Interfaces, 2014, 6, 244–251 CrossRef CAS PubMed.
  377. O. Nakhaei, N. Shahtahmassebi, M. R. Roknabadi and M. Behdani, Appl. Phys. A: Mater. Sci. Process., 2016, 122, 537 CrossRef.
  378. K. Nasouri and A. M. Shoushtari, Compos. Sci. Technol., 2017, 145, 46–54 CrossRef CAS.
  379. K. Nasouri, A. M. Shoushtari and M. R. M. Mojtahedi, J. Polym. Res., 2016, 23, 71 CrossRef.
  380. K. Nasouri, A. M. Shoushtari and M. R. M. Mojtahedi, Polym. Compos., 2017, 38, 2026–2034 CrossRef CAS.
  381. M. Chowdhury and G. Stylios, J. Mater. Sci., 2011, 46, 3378–3386 CrossRef CAS.
  382. S. B. Lee, S. J. Hong, J. W. Yi, S. K. Lee, Y. H. Choa and J. B. Kim, Met. Mater. Int., 2012, 18, 165–170 CrossRef CAS.
  383. Z. Zhu, Y. Xu, B. Qi, G. Zeng, P. Wu, G. Liu, W. Wang, F. Cui and Y. Sun, Environ. Sci.: Nano, 2017, 4, 302–306 RSC.
  384. Z.-d. Zhang, H. Wang, C. Qin, S. Chen, X. Ji, K. Sun, M. Chen, R.-h. Fan and X. Han, Mater. Des., 2016, 89, 543–548 CrossRef CAS.
  385. N. Cai, M. Chen, M. Liu, J. Wang, L. Shen, J. Wang, X. Feng and F. Yu, J. Mol. Liq., 2019, 289, 111060 CrossRef CAS.
  386. X. Meng and S. Dong, J. Alloys Compd., 2019, 810, 151806 CrossRef CAS.
  387. K.-Y. A. Lin, J.-T. Lin, X.-Y. Lu, C. Hung and Y.-F. Lin, J. Colloid Interface Sci., 2017, 505, 728–735 CrossRef CAS PubMed.
  388. H.-Y. Mi, X. Jing, H. Xie, H.-X. Huang and L.-S. Turng, Chem. Eng. J., 2018, 337, 541–551 CrossRef CAS.
  389. N. A. M. Barakat, B. Kim, S. J. Park, Y. Jo, M.-H. Jung and H. Y. Kim, J. Mater. Chem., 2009, 19, 7371–7378 RSC.
  390. N. A. M. Barakat, B. Kim, C. Yi, Y. Jo, M.-H. Jung, K. H. Chu and H. Y. Kim, J. Phys. Chem. C, 2009, 113, 19452–19457 CrossRef CAS.
  391. J. Yu, D.-H. Nam, Y.-J. Lee and J. Y. Chang, J. Microelectron. Electron. Packag., 2016, 23, 57–63 CrossRef.
  392. S.-R. Huang, K.-F. Lin, C.-F. Lee and W.-Y. Chiu, J. Polym. Sci., Part A: Polym. Chem., 2014, 52, 848–856 CrossRef CAS.
  393. D. Li, L. Luo, Z. Pang, L. Ding, Q. Wang, H. Ke, F. Huang and Q. Wei, ACS Appl. Mater. Interfaces, 2014, 6, 5144–5151 CrossRef CAS PubMed.
  394. T. Ren, Y. Si, J. Yang, B. Ding, X. Yang, F. Hong and J. Yu, J. Mater. Chem., 2012, 22, 15919–15927 RSC.
  395. R. K. Sadasivam, S. Mohiyuddin and G. Packirisamy, ACS Omega, 2017, 2, 6556–6569 CrossRef CAS PubMed.
  396. X. Huang, Y. Chen, J. Yu, J. Zhang, T. Sang, G. Tao and H. Zhu, J. Mater. Sci.: Mater. Electron., 2015, 26, 3474–3478 CrossRef CAS.
  397. Y. Yuan, Y. Feng, X. Zhao, M. Yang, H. Lu, J. Li, Y. Du, Y. Li and X. He, Ceram. Int., 2018, 44, 6673–6680 CrossRef CAS.
  398. Y. Si, T. Ren, Y. Li, B. Ding and J. Yu, Carbon, 2012, 50, 5176–5185 CrossRef CAS.
  399. J. Wu, N. Wang, Y. Zhao and L. Jiang, Nanoscale, 2015, 7, 2625–2632 RSC.
  400. Z. Ying, C. Z. Jing, Z. Jin, M. Z. Yong, F. Lin and J. Lei, ChemPhysChem, 2006, 7, 336–341 CrossRef PubMed.
  401. R. Guo, T. Jiao, R. Xing, Y. Chen, W. Guo, J. Zhou, L. Zhang and Q. Peng, Nanomaterials, 2017, 7, 317 CrossRef PubMed.
  402. N. A. M. Barakat, J. Mater. Sci., 2012, 47, 6237–6245 CrossRef CAS.
  403. F. Mou, J.-g. Guan, W. Shi, Z. Sun and S. Wang, Langmuir, 2010, 26, 15580–15585 CrossRef CAS PubMed.
  404. W. W. Pan, Q. F. Liu, R. Han, X. Chi and J. B. Wang, Appl. Phys. A: Mater. Sci. Process., 2013, 113, 755–761 CrossRef CAS.
  405. G. George and S. Anandhan, RSC Adv., 2014, 4, 62009–62020 RSC.
  406. H. K. Kim, W. Honda, B. S. Kim and I. S. Kim, J. Mater. Sci., 2013, 48, 1111–1116 CrossRef CAS.
  407. A. Garcia-Marquez, S. Glatzel, A. Kraupner, K. Kiefer, K. Siemensmeyer and C. Giordano, Chem. – Eur. J., 2018, 24, 4895–4901 CrossRef CAS PubMed.
  408. J. Zhu, S. Wei, D. Rutman, N. Haldolaarachchige, D. R. Young and Z. Guo, Polymer, 2011, 52, 2947–2955 CrossRef CAS.
  409. N. A. M. Barakat, M. F. Abadir, M. S. Akhtar, M. El-Newehy, Y.-S. Shin and H. Y. Kim, Mater. Lett., 2012, 85, 120–123 CrossRef CAS.
  410. A. Baranowska-Korczyc, A. Reszka, K. Sobczak, B. Sikora, P. Dziawa, M. Aleszkiewicz, L. Klopotowski, W. Paszkowicz, P. Dluzewski, B. J. Kowalski, T. A. Kowalewski, M. Sawicki, D. Elbaum and K. Fronc, J. Sol-Gel Sci. Technol., 2012, 61, 494–500 CrossRef CAS.
  411. X. Huang and S. W. Or, Mater. Lett., 2019, 238, 271–274 CrossRef CAS.
  412. R. Zhang, H. Wu, D. D. Lin and W. Pan, J. Am. Ceram. Soc., 2010, 93, 605–608 CrossRef CAS.
  413. P. Mohanapriya and N. V. Jaya, J. Nanosci. Nanotechnol., 2015, 15, 2226–2233 CrossRef CAS PubMed.
  414. X. Y. Li, J. N. Wang, L. L. Zhang and C. J. Li, J. Mater. Sci., 2012, 47, 465–472 CrossRef CAS.
  415. G. G. Philip, A. Senthamizhan, T. S. Natarajan, G. Chandrasekaran and H. A. Therese, Ceram. Int., 2015, 41, 13361–13365 CrossRef.
  416. J. Xiang, G. Z. Zhou, X. Q. Shen, Y. Q. Chu and Y. T. Guo, J. Sol-Gel Sci. Technol., 2012, 62, 186–192 CrossRef CAS.
  417. Z. Zhu, C. Ji, L. Zhong, S. Liu, F. Cui, H. Sun and W. Wang, J. Mater. Chem. A, 2017, 5, 18071–18080 RSC.
  418. Y. Hou, Y. Zhang, X. Du, Y. Yang, C. Deng, Z. Yang, L. Zheng and L. Cheng, RSC Adv., 2018, 8, 33574–33582 RSC.
  419. N. A. M. Barakat, K. A. Khalil, I. H. Mahmoud, M. A. Kanjwal, F. A. Sheikh and H. Y. Kim, J. Phys. Chem. C, 2010, 114, 15589–15593 CrossRef CAS.
  420. N. A. M. Barakat, K. A. Khalil and H. Y. Kim, Mater. Res. Bull., 2012, 47, 2140–2147 CrossRef CAS.
  421. K. R. Sapkota, P. Gyawali, A. Forbes, I. L. Pegg and J. Philip, J. Appl. Phys., 2012, 111, 123906 CrossRef.
  422. J. Lee, T.-Y. Hwang, M. K. Kang, G. Lee, H.-B. Cho, J. Kim and Y.-H. Choa, Appl. Surf. Sci., 2019, 471, 273–276 CrossRef CAS.
  423. N. A. M. Barakat, A. A. Elzatahry and K. A. Khalil, Mater. Res. Bull., 2014, 49, 503–508 CrossRef CAS.
  424. H. P. Li, X. L. Qin, W. Zhang and W. Pan, J. Am. Ceram. Soc., 2012, 95, 217–222 CrossRef CAS.
  425. C. J. Li and G. R. Xu, Mater. Sci. Eng., B, 2011, 176, 552–558 CrossRef CAS.
  426. P. Moradipour, L. Rajabi and F. Dabirian, J. Aust. Ceram. Soc., 2018, 54, 23–32 CrossRef CAS.
  427. Y. Luo, S. Nartker, M. Wiederoder, H. Miller, D. Hochhalter, L. T. Drzal and E. C. Alocilja, IEEE Trans. Nanotechnol., 2012, 11, 676–681 Search PubMed.
  428. W. Sun, X. Lu, Y. Xue, Y. Tong and C. Wang, Macromol. Mater. Eng., 2014, 299, 361–367 CrossRef CAS.
  429. W. Pan, Z. Ma, J. Liu, Q. Liu and J. Wang, Mater. Lett., 2011, 65, 3269–3271 CrossRef CAS.
  430. S. Nilmoung, P. Kidkhunthod, S. Pinitsoontorn, S. Rujirawat, R. Yimnirun and S. Maensiri, Appl. Phys. A: Mater. Sci. Process., 2015, 119, 141–154 CrossRef CAS.
  431. J. D. Starr, M. A. K. Budi and J. S. Andrew, J. Am. Ceram. Soc., 2015, 98, 12–19 CrossRef CAS.
  432. J. Fu, J. Zhang, Y. Peng, J. Zhao, G. Tan, N. J. Mellors, E. Xie and W. Han, Nanoscale, 2012, 4, 3932–3936 RSC.
  433. F. Bi, X. Dong, J. Wang and G. Liu, New J. Chem., 2015, 39, 3444–3451 RSC.
  434. Z. Zhang, G. Yang, J. Wei, H. Bian, J. Gao, J. Li and T. Wang, J. Cryst. Grow., 2016, 445, 42–46 CrossRef CAS.
  435. P. D. Prasad and J. Hemalatha, Mater. Res. Express, 2019, 6, 094007 CrossRef CAS.
  436. M. J. Bauer, X. Wen, P. Tiwari, D. P. Arnold and J. S. Andrew, Microsyst. Nanoeng., 2018, 4, 37 CrossRef PubMed.
  437. S.-H. Xie, J. Li, Y. Qiao, Y. Liu, L. Lan, Y. Zhou and S. Tan, Appl. Phys. Lett., 2008, 92, 062901 CrossRef.
  438. L. N. Pan, D. R. Cao, P. P. Jing, J. B. Wang and Q. F. Liu, Nanoscale Res. Lett., 2015, 10, 131 CrossRef PubMed.
  439. R. Lu, M. Xu, B. Fu, Y. Zhang, C. Zhou, Y. Zeng, S. Yang, X. Song and X. Zhou, ChemistrySelect, 2016, 1, 1510–1514 CrossRef CAS.
  440. J. Dai, C. Cheng, Z. Li and W. Feng, Mater. Res. Express, 2019, 6, 086116 CrossRef CAS.
  441. E. Santala, M. Kemell, M. Leskela and M. Ritala, Nanotechnology, 2009, 20, 035602 CrossRef PubMed.
  442. J. Zhang, J. Fu, G. Tan, F. Li, C. Luo, J. Zhao, E. Xie, D. Xue, H. Zhang, N. J. Mellors and Y. Peng, Nanoscale, 2012, 4, 2754–2759 RSC.
  443. D. Li, T. Herricks and Y. N. Xia, Appl. Phys. Lett., 2003, 83, 4586–4588 CrossRef CAS.
  444. J. Zhang, S. Zhu, J. Ming, L. Qiao, F. Li, A. Karim, Y. Peng and J. Fu, Nanoscale, 2019, 11, 13824–13831 RSC.
  445. J. Zhang, R. Kurosawa, H. Suematsu, T. Nakayama and S. S. Kim, Met. Mater. Int., 2012, 18, 505–508 CrossRef CAS.
  446. X. K. Zhang, J. Xiang, Z. P. Wu, L. Gong, X. Chen, G. G. Guan, Y. Wang and K. Y. Zhang, J. Alloys Compd., 2018, 764, 691–700 CrossRef CAS.
  447. Y. M. Ren, H. Y. Zhang, H. Z. An, Y. Zhao, J. Feng, L. L. Xue, T. Z. Luan and Z. J. Fan, J. Colloid Interface Sci., 2018, 526, 347–355 CrossRef CAS PubMed.
  448. H.-R. Jung, K. N. Kim and W.-J. Lee, Korean J. Chem. Eng., 2019, 36, 807–815 CrossRef CAS.
  449. P. Wang, T. Dong, M. Li and P. Yang, J. Nanosci. Nanotechnol., 2019, 19, 4474–4480 CrossRef CAS PubMed.
  450. J. Yan, S. Gao, C. Wang, B. Chai, J. Li, G. Song and S. Chen, Mater. Lett., 2016, 184, 43–46 CrossRef CAS.
  451. X. Huang, J. Zhang, S. Xiao, T. Sang and G. Chen, Mater. Lett., 2014, 124, 126–128 CrossRef CAS.
  452. C. Wang, X. Tan, J. Yan, B. Chai, J. Li and S. Chen, Appl. Surf. Sci., 2017, 396, 780–790 CrossRef CAS.
  453. J. Xie, Q. Wu and D. Zhao, Carbon, 2012, 50, 800–807 CrossRef CAS.
  454. M. Arias, V. M. Pantojas, O. Perales and W. Otano, J. Magn. Magn. Mater., 2011, 323, 2109–2114 CrossRef CAS PubMed.
  455. X. G. Huang, J. Zhang, S. R. Xiao and G. S. Chen, J. Am. Ceram. Soc., 2014, 97, 1363–1366 CrossRef CAS.
  456. J. Fu, J. Zhang, Y. Peng, C. Zhao, Y. He, Z. Zhang, X. Pan, N. J. Mellors and E. Xie, Nanoscale, 2013, 5, 12551–12557 RSC.
  457. Z. Hou, J. Xiang, X. Zhang, L. Gong, J. Mi, X. Shen and K. Zhang, J. Mater. Sci.: Mater. Electron., 2018, 29, 12258–12268 CrossRef CAS.
  458. J.-H. Nam, Y.-H. Joo, J.-H. Lee, J. H. Chang, J. H. Cho, M. P. Chun and B. I. Kim, J. Magn. Magn. Mater., 2009, 321, 1389–1392 CrossRef CAS.
  459. N. Ghazi, H. M. Chenari and F. E. Ghodsi, J. Magn. Magn. Mater., 2018, 468, 132–140 CrossRef CAS.
  460. M. Li, H.-Y. Bai, Z.-L. Da, X. Yan, C. Chen, J.-H. Jiang, W.-Q. Fan and W.-D. Shi, Cryst. Res. Technol., 2015, 50, 244–249 CrossRef CAS.
  461. C. J. Li, J. N. Wang, B. Wang, J. R. Gong and Z. Lin, J. Nanosci. Nanotechnol., 2012, 12, 2496–2502 CrossRef CAS PubMed.
  462. J.-M. Li and X.-L. Zeng, Appl. Phys. Lett., 2017, 110, 083107 CrossRef.
  463. A. Karaphun, S. Kaewmala, N. Meethong, S. Hunpratub and E. Swatsitang, J. Supercond. Novel Magn., 2018, 31, 1909–1916 CrossRef CAS.
  464. F. Song, X. Shen, M. Liu and J. Xiang, J. Solid State Chem., 2012, 185, 31–36 CrossRef CAS.
  465. F. Song, X. Shen, M. Liu and J. Xiang, J. Nanosci. Nanotechnol., 2011, 11, 6979–6985 CrossRef CAS PubMed.
  466. M. Q. Liu, F. Z. Song, X. Q. Shen and Y. W. Zhu, J. Sol-Gel Sci. Technol., 2010, 56, 39–46 CrossRef CAS.
  467. F.-Z. Mou, J.-G. Guan, Z.-G. Sun, X.-A. Fan and G.-X. Tong, J. Solid State Chem., 2010, 183, 736–743 CrossRef CAS.
  468. J. Liu, C. Gong and G. Fan, J. Sol-Gel Sci. Technol., 2012, 61, 185–191 CrossRef CAS.
  469. J. Zhang, J. Fu, F. Li, E. Xie, D. Xue, N. J. Mellors and Y. Peng, ACS Nano, 2012, 6, 2273–2280 CrossRef CAS PubMed.
  470. J. Xiang, Z. Hou, X. Zhang, L. Gong, Z. Wu and J. Mi, J. Alloys Compd., 2018, 737, 412–420 CrossRef CAS.
  471. J. S. Im, J. G. Kim, S.-H. Lee and Y.-S. Lee, Mater. Chem. Phys., 2010, 124, 434–438 CrossRef CAS.
  472. B. Li, C. Wang, W. Zhang, C. Hang, J. Fei and H. Wang, Mater. Lett., 2013, 91, 55–58 CrossRef CAS.
  473. D. Q. Zhang, J. Y. Cheng, J. X. Chai, J. J. Deng, R. Ren, Y. Su, H. Wang, C. Q. Ma, C. S. Lee, W. J. Zhang, G. P. Zheng and M. S. Cao, J. Alloys Compd., 2018, 740, 1067–1076 CrossRef CAS PubMed.
  474. X. Zhou, J. Xue, D. Zhou, Z. Wang, Y. Bai, X. Wu, X. Liu and J. Meng, ACS Appl. Mater. Interfaces, 2010, 2, 2689–2693 CrossRef CAS PubMed.
  475. J. Zheng, K. Du, D. Xiao, Z.-Y. Zhou, W.-G. Wei, J.-J. Chen, L.-F. Yin and J. Shen, Chin. Phys. Lett., 2016, 33, 097501 CrossRef.
  476. W. Ponhan, V. Amornkitbamrung and S. Maensiri, J. Alloys Compd., 2014, 606, 182–188 CrossRef CAS.
  477. M. Q. Liu, X. Q. Shen, F. Z. Song, J. Xiang and R. J. Liu, J. Sol-Gel Sci. Technol., 2011, 59, 553–560 CrossRef CAS.
  478. J. Dai, Y. Dai, Z. Wang and H. Gao, J. Exp. Nanosci., 2015, 10, 249–257 CrossRef CAS.
  479. P. Mohanapriya, R. Pradeepkumar, N. V. Jaya and T. S. Natarajan, Appl. Phys. Lett., 2014, 105, 022406 CrossRef.
  480. M. Liu, H. Liu, S. Sun, X. Li, Y. Zhou, Z. Hou and J. Lin, Langmuir, 2014, 30, 1176–1182 CrossRef CAS PubMed.
  481. Y. T. Wei, L. Zhong, D. Li, Q. L. Ma and X. T. Dong, Opt. Mater., 2019, 95, 109261 CrossRef CAS.
  482. D. Li, Q. L. Ma, Y. Song, X. Xi, X. T. Dong, W. S. Yu, J. X. Wang and G. X. Liu, RSC Adv., 2016, 6, 113045 RSC.
  483. D. Li, Q. L. Ma, X. Xi, X. T. Dong, W. S. Yu, J. X. Wang and G. X. Liu, Chem. Eng. J., 2017, 309, 230–239 CrossRef CAS.
  484. J. Chen, Y. Song, D. Li, Q. Ma, X. Dong, W. Yu, X. Wang, Y. Yang, J. Wang and G. Liu, J. Lumin., 2019, 206, 509–517 CrossRef CAS.
  485. R. Tao, C. Shao, X. Li, X. Li, S. Liu, S. Yang, C. Zhao and Y. Liu, J. Colloid Interface Sci., 2018, 529, 404–414 CrossRef CAS PubMed.
  486. M. Sobhan, Q. Xu, Q. Yang, F. Anariba and P. Wu, Appl. Phys. Lett., 2014, 104, 051606 CrossRef.
  487. S. H. Xie, J. Y. Li, R. Proksch, Y. M. Liu, Y. C. Zhou, Y. Y. Liu, Y. Ou, L. N. Lan and Y. Qiao, Appl. Phys. Lett., 2008, 93, 222904 CrossRef.
  488. S. You, C. Liu, H. Liu, X. Yu, S. Li, W. Liu, S. Guo and X. Zhao, Mater. Lett., 2014, 130, 157–159 CrossRef CAS.
  489. K. Prashanthi and T. Thundat, Scanning, 2014, 36, 224–230 CrossRef CAS PubMed.
  490. Y. W. Zhao, H. Q. Fan, G. C. Liu, Z. Y. Liu and X. H. Ren, J. Alloys Compd., 2016, 675, 441–447 CrossRef CAS.
  491. E. A. Duarte, P. A. Quintero, M. W. Meisel and J. C. Nino, Phys. C, 2013, 495, 109–113 CrossRef CAS.
  492. F. C. Lu, K. Yin, K. X. Fu, Y. N. Wang, J. Ren and Q. Xie, Ceram. Int., 2017, 43, 16101–16106 CrossRef CAS.
  493. B. Li, C. Wang, W. Liu, M. Ye and N. Wang, Mater. Lett., 2013, 90, 45–48 CrossRef CAS.
  494. T. Yang, J. Zheng, Q. Cheng, Y.-Y. Hu and C. K. Chan, ACS Appl. Mater. Interfaces, 2017, 9, 21773–21780 CrossRef CAS PubMed.
  495. X. Li, F. Guo, S. Y. Wang, X. Wang, X. L. Xu, J. Gao and W. F. Liu, Appl. Phys. Lett., 2015, 107, 062903 CrossRef.
  496. D. Baiyila, X. Wang, X. Li, B. Sharileaodu, X. Li, L. Xu, Z. Liu, L. Duan and J. Liu, J. Mater. Chem. A, 2014, 2, 12304–12310 RSC.
  497. V. D. M. Schiefferdecker, G. M. O. Barra, S. D. A. S. Ramoa and C. Merlini, Front. Mater., 2019, 6, 193 CrossRef.

Footnote

Y. Jia and C. Yang contributed equally to this work.

This journal is © The Royal Society of Chemistry 2021