Surface-modified Nanobiomaterials for Electrochemical and Biomedicine Applications [1st ed.] 9783030555016, 9783030555023

The series Topics in Current Chemistry Collections presents critical reviews from the journal Topics in Current Chemistr

324 38 15MB

English Pages IX, 255 [260] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Surface-modified Nanobiomaterials for Electrochemical and Biomedicine Applications [1st ed.]
 9783030555016, 9783030555023

Table of contents :
Front Matter ....Pages i-ix
TiO2 Nanomaterials in Photoelectrochemical and Electrochemiluminescent Biosensing (Xiangui Ma, Chao Wang, Fengxia Wu, Yiran Guan, Guobao Xu)....Pages 1-17
DNA-Iron oxide nanoparticles conjugates: functional magnetic nanoplatforms in biomedical applications (José Raúl Sosa‑Acosta, Claudia Iriarte‑Mesa, Greter A. Ortega, Alicia M. Díaz‑García)....Pages 19-47
Magnetic Nanoparticles as MRI Contrast Agents (Ashish Avasthi, Carlos Caro, Esther Pozo‑Torres, Manuel Pernia Leal, María Luisa García‑Martín)....Pages 49-91
Gold, silver and iron oxide nanoparticles: Synthesis and bionanoconjugation strategies aiming to electrochemical applications (Claudia Iriarte‑Mesa, Yeisy C. López, Yasser Matos‑Peralta, Karen de la Vega‑Hernández, Manuel Antuch)....Pages 93-132
Quantum Dot bioconjugates for diagnostic applications (María Díaz‑González, Alfredo de la Escosura‑Muñiz, Maria Teresa Fernandez‑Argüelles, Francisco Javier García Alonso, Jose Manuel Costa‑Fernandez)....Pages 133-176
Carbon Nanotubes in Biomedicine (Viviana Negri, Jesús Pacheco‑Torres, Daniel Calle, Pilar López‑Larrubia)....Pages 177-217
Bioconjugated Plasmonic Nanoparticles for Enhanced Skin Penetration (David Alba‑Molina, Juan J. Giner‑Casares, Manuel Cano)....Pages 219-235
Proteins-based nanocatalyts for energy conversion reactions (Daily Rodriguez‑Padron, Md Ariful Ahsan, Mohamed Fathi Sanad, Rafael Luque, Alain R. Puente Santiago)....Pages 237-255

Citation preview

Topics in Current Chemistry Collections

Alain R. Puente-Santiago Daily Rodríguez-Padrón   Editors

Surface-modified Nanobiomaterials for Electrochemical and Biomedicine Applications

Topics in Current Chemistry Collections

Journal Editors Massimo Olivucci, Siena, Italy and Bowling Green, USA Wai-Yeung Wong, Hong Kong, China Series Editors Hagan Bayley, Oxford, UK Greg Hughes, Codexis Inc, USA Christopher A. Hunter, Cambridge, UK Seong-Ju Hwang, Seoul, South Korea Kazuaki Ishihara, Nagoya, Japan Barbara Kirchner, Bonn, Germany Michael J. Krische, Austin, USA Delmar Larsen, Davis, USA Jean-Marie Lehn, Strasbourg, France Rafael Luque, Córdoba, Spain Jay S. Siegel, Tianjin, China Joachim Thiem, Hamburg, Germany Margherita Venturi, Bologna, Italy Chi-Huey Wong, Taipei, Taiwan Henry N.C. Wong, Hong Kong, China Vivian Wing-Wah Yam, Hong Kong, China Chunhua Yan, Beijing, China Shu-Li You, Shanghai, China

Aims and Scope The series Topics in Current Chemistry Collections presents critical reviews from the journal Topics in Current Chemistry organized in topical volumes. The scope of coverage is all areas of chemical science including the interfaces with related disciplines such as biology, medicine and materials science. The goal of each thematic volume is to give the non-specialist reader, whether in academia or industry, a comprehensive insight into an area where new research is emerging which is of interest to a larger scientific audience. Each review within the volume critically surveys one aspect of that topic and places it within the context of the volume as a whole. The most significant developments of the last 5 to 10 years are presented using selected examples to illustrate the principles discussed. The coverage is not intended to be an exhaustive summary of the field or include large quantities of data, but should rather be conceptual, concentrating on the methodological thinking that will allow the non-specialist reader to understand the information presented. Contributions also offer an outlook on potential future developments in the field. More information about this series at http://www.springer.com/series/14181

Alain R. Puente-Santiago Daily Rodríguez-Padrón Editors

Surface-modified Nanobiomaterials for Electrochemical and Biomedicine Applications With contributions from Md Ariful Ahsan • David Alba‑Molina • Francisco Javier García Alonso Manuel Antuch • Ashish Avasthi • Daniel Calle • Manuel Cano Carlos Caro • Jose Manuel Costa‑Fernandez • Alicia M. Díaz‑García María Díaz‑González • Alfredo de la Escosura‑Muñiz Maria Teresa Fernandez‑Argüelles • María Luisa García‑Martín Juan J. Giner‑Casares • Yiran Guan • Claudia Iriarte‑Mesa Manuel Pernia Leal • Yeisy C. López • Pilar López‑Larrubia Rafael Luque • Xiangui Ma • Yasser Matos‑Peralta • Viviana Negri Greter A. Ortega • Jesús Pacheco‑Torres • Esther Pozo‑Torres Daily Rodriguez‑Padron • Mohamed Fathi Sanad Alain R. Puente Santiago • José Raúl Sosa‑Acosta Karen de la Vega‑Hernández • Chao Wang • Fengxia Wu Guobao Xu

Editors Alain R. Puente-Santiago Department of Chemistry The University of Texas at El Paso El Paso, TX, USA

Daily Rodríguez-Padrón Department of Organic Chemistry University of Cordoba Córdoba, Córdoba, Spain

Partly previously published in Top Curr Chem (Z) Volume 378 (2020). ISSN 2367-4067 Topics in Current Chemistry Collections ISBN 978-3-030-55501-6 © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Contents

Preface .............................................................................................................. TiO2 Nanomaterials in Photoelectrochemical and Electrochemiluminescent Biosensing ............................................................. Xiangui Ma, Chao Wang, Fengxia Wu, Yiran Guan and Guobao Xu: Top Curr Chem (Z) 2020, 378:28 (3, March 2020) https://doi.org/10.1007/s41061-020-0291-y DNA–Iron Oxide Nanoparticles Conjugates: Functional Magnetic Nanoplatforms in Biomedical Applications ................................................... José Raúl Sosa‑Acosta, Claudia Iriarte‑Mesa, Greter A. Ortega and Alicia M. Díaz‑García: Top Curr Chem (Z) 2020, 378:13 (10, January 2020) https://doi.org/10.1007/s41061-019-0277-9 Magnetic Nanoparticles as MRI Contrast Agents ........................................ Ashish Avasthi, Carlos Caro, Esther Pozo‑Torres, Manuel Pernia Leal and María Luisa García‑Martín: Top Curr Chem (Z) 2020, 378:40 (7, May 2020) https://doi.org/10.1007/s41061-020-00302-w Gold, Silver and Iron Oxide Nanoparticles: Synthesis and Bionanoconjugation Strategies Aimed at Electrochemical Applications...................................................................................................... Claudia Iriarte‑Mesa,·Yeisy C. López, Yasser Matos‑Peralta, Karen de la Vega‑Hernández and Manuel Antuch: Top Curr Chem (Z) 2020, 378:12 (7, January 2020) https://doi.org/10.1007/s41061-019-0275-y Quantum Dot Bioconjugates for Diagnostic Applications ........................... María Díaz‑González, Alfredo de la Escosura‑Muñiz, Maria Teresa Fernandez‑Argüelles, Francisco Javier García Alonso and Jose Manuel Costa‑Fernandez: Top Curr Chem (Z) 2020, 378:35 (26, March 2020) https://doi.org/10.1007/s41061-020-0296-6

v

vii 1

19

49

93

133

vi

Contents

Carbon Nanotubes in Biomedicine ................................................................. Viviana Negri, Jesús Pacheco‑Torres, Daniel Calle and Pilar López‑Larrubia: Top Curr Chem (Z) 2020, 378:15 (14, January 2020) https://doi.org/10.1007/s41061-019-0278-8 Bioconjugated Plasmonic Nanoparticles for Enhanced Skin Penetration .............................................................................................. David Alba‑Molina, Juan J. Giner‑Casares and Manuel Cano: Top Curr Chem (Z) 2020, 378:8 (16, December 2019) https://doi.org/10.1007/s41061-019-0273-0 Proteins‑Based Nanocatalysts for Energy Conversion Reactions ............... Daily Rodriguez‑Padron, Md Ariful Ahsan, Mohamed Fathi Sanad, Rafael Luque and Alain R. Puente Santiago: Top Curr Chem (Z) 2020, 378:43 (10, June 2020) https://doi.org/10.1007/s41061-020-00306-6

177

219

237

Preface

The recent advances in the nanoscience field have opened new horizons towards the development of exciting nanoscale materials with promising applications.[1] Surface-decorated nanoarchitectures, from functionalized 0D-quantum dots to 3Dprotein nanohybrids, have been successfully used as functional materials for sensing, energy conversion and theranostic. [2] [3] [4] In this topical collection, the development of surface-functionalized nanobiosystems for electrochemical and biomedical applications is comprehensively addressed. The functionalization strategies of low-dimensional interfaces with biological structures such as DNA, proteins or genes to construct bioinorganic heterostructures, are carefully discussed. Furthermore, seminal examples correlate the structural effects of the interfacial functionalization on the improved catalytic, sensing and biomedical properties of the resulting nanobioassemblies, thus providing valuable information for the material science, biotechnology and catalysis communities. In this regard, Guobao Xu and coworkers summarize the most recent developments as well as the future prospects in photoelectrochemical and electrochemiluminescent biosensing using TiO2 nanocomposites. Alicia Garcia et al. discuss how the interfacial interactions of iron oxide nanoparticles (IONPs) and DNA molecules can facilitate the development of nanocarriers and gene delivered vectors as efficient magnetic hybrids. In this direction, Maria Luisa, Manuel Antuch and their coworkers, extensively review the use of IONPs-based biomaterials for MRI and electrochemical applications, respectively. Jose Manuel and collaborators elegantly describe the crucial role of the surface functionalization of quantum dots (QDs) to develop highly stable colloidal nanosupensions for biomolecule targeting, luminescent imaging and drug delivery. The contribution of Pilar Lopez et. al. is particularly interesting due to the analysis of the interaction of carbon nanotubes (CNTs) with living cells and their implication on the proteome and genome. Also, Manuel Cano and colleagues highlight the bioconjugation of plasmonic nanoparticles with different biological materials to fabricate efficient biocompatible materials for skin penetration. Finally Alain R. and collaborators address the role of the synergistic interactions between different types of proteins and low-dimensional nanomaterials to assemble superior bioelectrocatalytic nanosystems.

vii

viii

Preface

In summary, this topical collection provides in-depth information about the interfacial interactions that govern the assembly of diverse types of nanobiomaterials as well as their implications for electrochemical and biomedical applications. We believe this collection will motivate many researchers towards the development of the next generation of bio-inspired materials.

Alain R. Puente-Santiago Department of Chemistry, University of Texas at El Paso, El Paso, USA

Daily Rodríguez-Padrón Department of Organic Chemistry, University of Cordoba, Cordoba, Spain

Preface

ix

References 1. Jeong GH, Sasikala SP, Yun T, Lee GY, Lee WJ, Kim SO (2020) Adv. Mater.: 1907006. doi:10.1002/adma.201907006 2. Patel KD, Singh RK, Kim HW (2019) Mater. Horizons 6: 434. doi:10.1039/ c8mh00966j 3. Yan YB, Zhai D, Liu Y, Gong J, Chen J, Zan P, Zeng ZP, Li SZ, Huang W, Chen P (2020) ACS Nano 14: 1185. doi:10.1021/acsnano.9b09554 4. Le Goff A, Holzinger M (2018) Sustain. Energ. Fuels 2: 2555. doi:10.1039/ c8se00374b

Topics in Current Chemistry (2020) 378:28 https://doi.org/10.1007/s41061-020-0291-y REVIEW

TiO2 Nanomaterials in Photoelectrochemical and Electrochemiluminescent Biosensing Xiangui Ma1,2 · Chao Wang1,3 · Fengxia Wu1 · Yiran Guan1 · Guobao Xu1,2  Received: 30 December 2019 / Accepted: 14 February 2020 / Published online: 3 March 2020 © Springer Nature Switzerland AG 2020

Abstract Titanium dioxide ­ (TiO2) is increasingly being used in biosensing applications. Herein, we review the most recent developments in photoelectrochemical (PEC) and electrochemiluminescent (ECL) biosensing based on ­TiO2 nanomaterials, as well as the mechanisms that lead to the improved performance of biosensors that incorporate these nanomaterials. The merits of ­TiO2-based ECL and PEC biosensing strategies are summarized by highlighting some illustrative examples that have been reported within the last 5 years. The future prospects for and challenges in this field are also discussed. Keywords TiO2 · Biosensing · Photoelectrochemistry · Electrochemiluminescence · Nanomaterials

1 Introduction Photoelectrochemical biosensing is promising novel detection method that focuses on photoelectron transfer processes at biomodified electrode/solution interfaces. The resulting photocurrent reflects the production of an electron donor/acceptor in

Chapter 1 was originally published as Ma, X., Wang, C., Wu, F., Guan, Y. & Xu, G. Topics in Current Chemistry (2020) 378: 28. https://doi.org/10.1007/s41061-020-0291-y. * Yiran Guan [email protected] * Guobao Xu [email protected] 1

State Key Laboratory of Electroanalytical Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, Jilin, People’s Republic of China

2

University of Science and Technology of China, Hefei 230026, Anhui, People’s Republic of China

3

College of Chemistry and Bioengineering, Guilin University of Technology, Guilin 541004, People’s Republic of China



Reprinted from the journal

1

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:28

a target-recognition process when it is irradiated by light of a suitable wavelength. PEC sensing analysis is more sensitive and has a lower background than traditional electrochemical and optical detection techniques because it separates the excitation source (light) from the output signal (photocurrent). Because of its low cost, high sensitivity, very low background noise, short response time, ease of use, and ability to be miniaturized, PEC biosensing has rapidly been applied in various contexts, including for food detection, environmental monitoring, gene testing, and early clinical diagnosis [1–3]. Electrochemiluminescence (ECL)—electrogenerated chemiluminescence—can be considered the reverse of PEC. ECL is the electrochemical process in which an excited species is electrogenerated at an electrode surface and then allowed to relax to a less excited state, resulting in the emission of light. As it is a combined electrochemical and spectroscopic technique, ECL permits both spectral and temporospatial resolution, which makes it a powerful tool for both sensing and imaging [4, 5]. The literature on ECL biosensing methods is extensive, which reflects their inherent advantages, including rapidity, simplicity, high sensitivity, low cost, and nearzero background signal. The integration of nanotechnology into ECL biosensors has also significantly enhanced the sensitivity and diversity of ECL biosensors in recent years [6]. TiO2 nanomaterials are commonly used in PEC and ECL biosensors due to their advantageous properties, such as strong light absorption, chemical and mechanical stability, good catalytic ability, high biocompatibility, and large specific surface areas [7, 8]. There are three crystalline forms of ­TiO2: anatase, rutile, and brookite. The utilization of T ­ iO2 nanomorphologies such as nanosheets [9], nanopillars [10], nanoarrays [11], nanoparticles [12], nanorods [13], nanowires [14], nanoneedles [15, 16], nanoflowers [17], nanocubes [18], and mesocrystals [19, 20] in many advanced bioanalytical strategies is frequently reported. In this mini review, we present recent advances in PEC and ECL biosensing based on ­TiO2 nanomaterials, exploring in detail the sensing strategies used and response mechanisms involved. However, given the limited space available in a mini review, we concentrate on the most significant aspects of and important advances in this field.

2 Applications of ­TiO2 Nanomaterials in Photoelectrochemistry Biosensing In PEC biosensing, the photoactive material significantly influences the analytical performance. However, a well-known disadvantage of ­TiO2-based photoelectrochemistry is that the wide band gap of ­TiO2 results in peak photoabsorption in the ultraviolet (UV; λ ≤ 387 nm) [14, 21]. UV light decreases biomolecular activity, which implies that the applicability of T ­ iO2 to PEC bioanalysis is limited [22]. Low surface charge-transfer efficiency and high electron–hole pair recombination also limit the sensitivities of T ­ iO2-based biosensors. Therefore, attempts have been made to identify new photoactive forms of ­TiO2 that could be used to develop improved biosensors. Countless works in the fields of materials and physical chemistry have

13

2

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28

provided us with the theoretical basis for developing ­ TiO2 nanomaterials with enhanced properties [3, 8]. In this section, the rational design and engineering of various ­TiO2 nanomaterials to minimize the shortcomings of pure ­TiO2 are discussed. Many methods have been used to obtain high-performance PEC biosensors based on ­TiO2 nanomaterials, including the coupling of T ­ iO2 with a narrow bandgap semiconductor, the dye sensitization of ­TiO2, the deposition of noble metal nanoparticles onto ­TiO2, the doping of ­TiO2 with metal and nonmetal atoms, and the preparation of ­TiO2 with engineered defects. Due to the ability to control their band gap and size, quantum dots (QDs) are commonly combined with other photoactive materials to achieve improved PEC behavior [23]. Chen et  al. [9] reported a PEC immunoassay strategy (see Fig.  1a) utilizing ­TiO2 nanosheets and CdS QDs. Due to the narrower band gap of the CdS QDs (2.4 eV) than that of the anatase ­TiO2 nanosheets (3.2 eV), the QDs effectively sensitized the ­TiO2 nanosheets to visible light. The ­TiO2/CdS QDs composite produced strong and stable photoelectric signals and showed strong and broad absorption in the visible light region. This biocompatible nanocomposite also provided

Fig. 1  a Schematic of electron transfer between a CdS-sensitized ­TiO2 nanosheet electrode and Ag@ Cu2O–Ab2 composites used as a label in the PEC detection of cTnI. b Schematic of the PEC mechanism for ­MgIn2S4–TiO2 heterojunction-based aptasensing of ATP. c Schematic of a PEC bioassay for CEA that utilizes porphyrin-sensitized ­TiO2. Reproduced with permission from [9, 11, 12] Reprinted from the journal

3

13



Topics in Current Chemistry (2020) 378:28

numerous functional groups at which to immobilize antibodies ­(Ab1) for cTnI (cardiac troponin I). In another study, mesoporous T ­ iO2 functionalized with CdS QDs (acting as a sensitizer) was synthesized to facilitate enhanced PEC aptasensing of PSA (prostate-specific antigen) [24]. Ordered mesoporous T ­ iO2 is an ideal platform for incorporating CdS QDs due to its multiple scattering properties and large surface area. The combination of QDs and mesoporous ­TiO2 was found to promote charge transfer and electron–hole pair separation at the interface. Recently, graphite-like carbon nitride (g-C3N4) QDs and N-doped graphene QDs (N-GQDs) [10] were used to sensitize ­TiO2 nanopillars and therefore achieve stronger visible light absorbance for the sensitive PEC detection of pcDNA3-HBV (hepatitis B virus). The band gaps of the T ­ iO2, the g-C3N4 QDs, and the N-GQDs were 3.2, 2.76, and 1.7 eV, respectively. After incorporating the g-C3N4 and N-GQDs, the charge transfer and electron–hole separation efficiency improved considerably due to photosensitization and the formation of a heterojunction between g-C3N4 and ­TiO2. The utilization of a heterostructure comprising ­TiO2 and a semiconductor with a suitable band gap and matched energy levels can tremendously decrease charge recombination, enhance the electron transfer efficiency, and increase light absorption in PEC biosensors. Liu et al. [13] presented a PEC biosensing strategy based on a heterostructure containing ­MoS2 nanosheets and ­TiO2 nanorods. The ­TiO2 nanorods served as hydrothermal growth templates for the ­MoS2 nanosheets, which were a few layers thick. The band edge of the ­MoS2 nanosheets was well matched to that of ­TiO2, facilitating the separation and transfer of the photogenerated charge. Due to the excellent biocompatibility of this system, glucose oxidase (GOx) was immobilized onto a ­MoS2/TiO2 nanocomposite-modified ITO electrode for the sensitive detection of glucose under visible light irradiation. In addition, a T ­ iO2–BiVO4 heterostructure [21] for PEC biosensing has been obtained by depositing B ­ iVO4 NPs onto ­TiO2 nanospheres using a solvothermal method. The narrow energy gap (2.34 eV) of ­BiVO4 was found to improve the absorption of the heterostructure in the visible light region. The high surface area and biocompatible microenvironment of the ­TiO2–BiVO4 heterostructure permitted improved loading of recognition biomolecules for 17β-estradiol measurements. Also, as shown in Fig.  1b, a heterojunction consisting of M ­ gIn2S4 nanoplates and a T ­ iO2 nanoarray was developed [11] for the PEC aptasensing of adenosine triphosphate (ATP). Interestingly, the ­TiO2 nanoarray was observed to have better photoelectric properties than T ­ iO2 NPs, including a narrower bandgap (~ 3.0 eV), higher conductivity, and reduced recombination of electron–hole pairs. The presence of the ­MgIn2S4/TiO2 nanoarray heterojunction improved visible light absorption, resulting in a ~ 6.8-fold increase in photocurrent compared with a more conventional ­TiO2 electrode. Ferrocene, an electron donor employed as a label, has been used to accelerate electron–hole separation and thus enhance the photocurrent in PEC biosensors. In particular, Wei’s group [15] reported the cytosensing of RAW264.7 macrophage cells utilizing a ­TiO2 nanoneedles@MoO3 array p–n heterojunction. ­MoO3 is a p-type semiconductor with a relatively wide band gap of 2.9 eV; its absorption peak is near to the UV light region. Coupling ­MoO3 with ­TiO2 yielded a p–n heterojunction that promoted visible light absorption due to the offset of 2.61 eV between the valence band (VB) of ­TiO2 and the conduction band (CB) of M ­ oO3. Immobilizing F4/80 antibodies used

13

4

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28

as recognition molecules onto this system allowed RAW264.7 cells to be measured directly by steric hindrance. Another common strategy for enhancing PEC biosensing is dye sensitization. For example, Tang’s group [12] reported an ultrasensitive PEC immunoassay of CEA based on porphyrin-sensitized T ­ iO2 nanoparticles (NPs) (Fig.  1c). Linking the water-soluble species 5,10,15,20-tetra(4-sulfophenyl)-21H,23H-porphyrin (TSPP) through its sulfo groups to the ­TiO2 led to an improvement in photoelectron transfer. Compared with T ­ iO2, TSPP has a narrower band gap (2.8 eV) and absorbs more strongly in the visible region due to its delocalized π-electron system. Closely matched energy levels of T ­ iO2 and TSPP result in fast electron transfer and slow charge recombination. The resulting system was then used in combination with glucose oxidase (GOx)-labeled ­Ab2 to fabricate a stable and sensitive PEC biosensor for the detection of CEA. When glucose was present, the GOx oxidized it to ­H2O2, which scavenged photogenerated holes in ­TiO2 at low potential, amplifying the photocurrent. Liu et al. [19] demonstrated that adding a chelating assembly of polydopamine (PDA) to the surfaces of rutile ­TiO2 mesocrystals enhanced the PEC performance of an immunoassay for zearalenone. The benzoquinone groups of the PDA received photoelectrons from the ­TiO2 mesocrystals and enhanced the photocathodic current. Due to its special chemical structure, PDA absorbs long-wavelength light, improving charge-carrier separation. Similarly, polymerized l-DOPA (PD) [25] was applied to improve the enzymatic performance of a PEC glucose sensor by coupling the PD to a core–shell gold nanorod@TiO2 heterostructure. The PD not only enhanced the light absorption of the PEC system but it also provided a biocompatible matrix for surface functionalization and biointeractions. The core–shell heterostructure was found to efficiently assist interfacial charge transfer. The photoactive current density of PD/AuNR@TiO2/FTO was about 8.4 times that of ­TiO2/FTO and 2.6 times that of AuNR@TiO2/FTO. Immobilizing GOx and HRP (horseradish peroxidase) on the composite yielded a sensitive enzymatic PEC sensor for glucose with a low detection limit. Yan [26] reported a turn-on PEC strategy based on localized surface plasmon resonance (LSPR) enhancement and dye sensitization for detecting the activity of protein kinase A (PKA) under visible light ­ iO2/ITO was irradiation. In the presence of PKA and ATP, the kemptide on the T phosphorylated and then linked with DNA-conjugated gold nanoparticle (AuNP) probes. [Ru(bpy)3]2+ was intercalated into the DNA grooves, where it could absorb visible light and generate photoexcited electrons and thus photocurrent under visible light irradiation. Meanwhile, the AuNPs were able to load a considerable amount of [Ru(bpy)3]2+-intercalated DNA, which enhanced the photocurrent transfer efficiency through LSPR. This phenomenon is the collective oscillation of electron clouds in highly conductive metal nanoparticles under suitable light irradiation. For nanoparticles of noble metals such as Ag and Au  [27], this oscillatory resonance occurs at visible wavelengths. Thus, these nanoparticles can facilitate electron transfer and electron–hole pair seperation via the LSPR effect. In 2014, Da et al. [14] reported a PEC biosensing strategy that utilized ­TiO2 nanowires decorated with Au NPs and was based on surface plasmon resonance (Fig. 2a). In this method, Au NPs were attached directly to T ­ iO2 nanowires, which led to double the photocurrent density compared Reprinted from the journal

5

13



Topics in Current Chemistry (2020) 378:28

Fig. 2  a Schematic of a T ­ iO2 PEC biosensor decorated with Au NPs. b Schematic of PEC detection based on Cu-doped ­TiO2. Reproduced with permission from [14, 30]

to that of pure T ­ iO2. Cholera toxin subunit B, which was used as a test analyte, was sensitively detected in real time using the proposed PEC strategy. Ju’s group [28] prepared a ternary compound consisting of ­TiO2 nanotubes, polyaniline (PANI), and gold NPs for a novel PEC bioassay based on LSPR. The LSPR band of the Au NPs improved the system’s capacity to absorb light at ~ 540 nm and enhanced its photocurrent transfer efficiency. The electrochromism of PANI enhanced the ability of the system to harvest visible light and to separate the charge. The biocompatibility of the ternary composite permitted the immobilization of lactate dehydrogenase (LDH) and ­NAD+ on the electrode, facilitating l-lactate PEC sensing with a detection limit of 0.15 μM. Hao et al. [29] synthesized Ag/TiO2-decorated, 3D-nitrogendoped graphene hydrogel (3DNGH) to promote PEC performance. The 3DNGH had a porous structure and a large surface area that could accommodate a considerable amount of Ag and ­TiO2 NPs. Under light irradiation, electrons on the Ag surface were transferred to T ­ iO2. This Ag/TiO2/3DNGH nanocomposite exhibited a photocurrent that was approximately 60 times greater than that afforded by pristine T ­ iO2 NPs, and a sensitive label-free PEC thrombin aptasensor was constructed from this nanocomposite. The rational design of metal- and nonmetal-doped ­TiO2 can reduce its band gap and improve its response to visible light. Tang et al. [18] reported a PEC strategy for aflatoxin B1 (AFB1) detection based on Ce-doped T ­ iO2 nanocube@MoSe2 nanosheets. Ce was doped into the T ­ iO2 nanocubes using a one-step hydrothermal method. The CB of the Ce-doped ­TiO2 is at a lower energy than that of of ­TiO2, so it has a smaller energy gap. ­MoSe2 nanosheets were then grown on the ­TiO2 nanocubes, yielding a system with a large surface area, and the resulting heterojunction composite presented enhanced visible light absorption. Finally, the AuNPs were introduced as a quenching label that markedly decreased the photocurrent, yielding an ultrasensitive assay for AFB1. Wei’s group [30] developed a Cu-doped ­TiO2/g-C3N4 PEC immunosensor for CEA (Fig. 2b). Doping Cu into ­TiO2 to give a Cu:TiO2 nanocomposite caused the band gap energy to shrink to 2.85 eV, leading to much stronger light absorption and significantly enhanced photocurrent. The PEC performance was further improved upon the addition of g-C3N4 due to the resulting photosensitation and a synergistic effect. Alkaline phosphatase (ALP) was employed

13

6

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28



as a label for A ­ b2 in order to catalyze the hydrolysis of ascorbic acid 2-phosphate to ascorbic acid, which acted as an electron donor, leading to even more detection sensitivity for CEA. Zhang et al. [31] presented a PEC aptasensor for the supersensitive detection of CEA based on nonmetal-doped ­TiO2 and CdS QDs used as a sensitizer. In that study, Br and N were codoped into T ­ iO2, which narrowed its band gap from 3.2  eV to 2.88  eV. The band gap of the (Br, N)-codoped T ­ iO2 was well matched to that of CdS QDs (2.4  eV), meaning that these QDs could be used to intensify the photocurrent. Exonuclease III (Exo-III)-assisted cycling was also applied in this strategy to achieve the supersensitive detection of CEA. The physicochemical properties of ­TiO2 semiconductors also depend on their intrinsic defects and extrinsic impurities. Tang’s group [32] developed ­TiO2 with engineered defects, ­TiO2–x ­(dTiO2–x), and modified the ­dTiO2–x with Au NPs to create a novel photoelectric material for sensitive PEC biosensing (Fig.  3). Oxygen vacancies were induced in the d­ TiO2–x by doping it with F ­ e3+, which narrowed the band gap (to 2.5 eV), extended the absorption edge, and intensified the visible light absorption. When the defective ­TiO2 was irradiated with 580-nm light, the photocurrent was found to be as much as 6.7-fold higher in the presence of Au NPs than in their absence due to the hot electron transfer facilitated by the LSPR of the Au NPs. A significantly smaller (only 2.4-fold) jump in photocurrent in the presence of Au NPs was observed for pristine ­TiO2 under the same conditions. As a result, the Au NPs acted as a photocurrent-enhancing label that facilitated sensitive PEC

Fig. 3  Schematics of PEC biosensor strategies based on the application of d­ TiO2–x (a) and pristine T ­ iO2 (b). c Effects of different excitation wavelengths on the photocurrent intensity. Reproduced with permission from [32] Reprinted from the journal

7

13



Topics in Current Chemistry (2020) 378:28

DNA detection. Guided by density functional theory (DFT) calculations, Fu et  al. [22] developed plasmonic Au-modified, bulk/surface defect-engineered ­TiO2 nanotube photonic crystals (Au/bsDE-TiO2 NTPCs) for the in  vivo near-infrared PEC aptasensing of tetracycline (TET). Bulk defects were introduced into the ­TiO2 lattice by high-temperature annealing in a reducing atmosphere or vacuum, and then Au NPs were sputtered onto the defective T ­ iO2 surface. The PEC response of the resulting Au/bsDE-TiO2 NTPCs nanocomposite extended into the near-infrared region (900  nm). The photocurrent of the proposed PEC system increased significantly when TET was captured by an aptamer on the nanocomposite. A thin (0.1 mm diameter) Ti wire modified with Au/bsDE-TiO2 NTPCs was then successfully used to monitor the TET in a mouse tail in vivo under near-infrared (NIR) light. Speaking of NIR PEC detection, Qiu et  al. [33] reported the application of core–shell ­NaYF4:Yb,Tm@TiO2 upconversion microrods for the detection of CEA (carcinoembryonic antigen). The ­Yb3+ ions acted as a photosensitizer, absorbing the near-infrared light at 980  nm and generating two emission peaks at 453 and 479 nm. At the same time, the doped ­Tm3+ emitted UV light (with emission peaks at 291, 348, and 363 nm) that overlapped closely with the absorption peak of ­TiO2. After modifying the the ­NaYF4:Yb,Tm with ­TiO2, the peak intensity of the photoluminescence of the resulting system in the UV region markedly decreased. The proposed PEC system exhibited a good response to the guanine bases generated during CEA aptasensing. Due to its low phototoxicity in biological systems, this ­NaYF4:Yb,Tm@TiO2-based biosensor should expand the application of upconversion materials to PEC detection.

3 Applications of ­TiO2 Nanomaterials in Electrochemiluminescence Biosensing Recently, ­TiO2 nanomaterials and their composites have drawn considerable interest from those working in the field of ECL biosensing, given the biocompatibility, large surface area, and unique ECL properties of these materials. ­TiO2 nanomaterials can be applied in a variety of roles in ECL biosensors, including as a matrix for biomolecules, as an ECL luminophore, and as a catalyst for the ECL reaction. In this section, ECL bioassays based on T ­ iO2 nanomaterials are explored via a number of illustrative examples. TiO2 is an ideal matrix for immobilizing biomolecules on an electrode. For instance, Dai and coauthors [20] presented a dual-signal ECL biosensor based on ­TiO2 mesocrystals and CdTe QDs for the detection of metallothioneins. A considerable amount of Ru(bpy)2+ TiO2 mesocrystal struc3 was immbolized in the porous ­ ture through ion exchange. In addition, the presence of ­TiO2 mesocrystals on the nanocomposite surface resulted in a higher pH, thus increasing the ECL intensity of Ru(bpy)2+ 3 . ­TiO2 is able to link to metallothioneins efficiently via strong interactions with the sulfhydryl groups of metallothioneins, which blocks electron transfer and hinders the diffusion of coreactants, thus decreasing the ECL intensity of Ru(bpy)2+ 3 . The CdTe QDs can link to the remaining SH groups of metallothioneins, generating a cathodic ECL. Therefore, a ratiometric biosensor was constructed for

13

8

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28



metallothioneins, which presented a LOD that reached down to the ng/mL level. Interestingly, Tsuneyasu et al. [34] proved that T ­ iO2 NPs amplify the ECL intensity of Ru(bpy)2+ , thus improving the properties of an ECL device based on Ru(bpy)2+ 3 3 . The optical and electrochemical properties of the resulting system were studied in detail, and it was found that the ECL improvement induced by the ­TiO2 NPs was due to their ability to suppress the nonradiative quenching of the excited states of Ru(bpy)2+ 3 . In another study, a dual-amplified ECL sensor based on tetragonal rutile ­TiO2 mesocrystals (TRM) was developed for sensitive zearalenone (ZEA) detection [35]. As shown in Fig. 4, the nanoporous TRM were functionalized with poly(amidoamine) dendrimers (PAAD) in order to generate a biocompatible substrate with a surface area large enough to immobilize sufficient A ­ b1 of ZEA. On the other hand, the TRM were employed to adsorb large amounts of Ru(bpy)2+ 3 and to immobilize ­Ab2. A sensitive sandwich-type ECL strategy was realized through these two TRM-based amplification techniques. Electron transfer from the reduced species of Ru(bpy)2+ 3 and extensive light scattering by TRM led to strong ECL emission and the sensitive detection of ZEA. In another approach to improving ECL performance, ­TiO2 MOFs [36] were synthesized via the calcination of MIL-125 (Tibased MOFs). These ­TiO2 MOFs had a highly regular pore structure and a large specific surface area, allowing the immobilization of large amounts of Ru(bpy)2+ 3 , β-cyclodextrin (β-CD), and antibody. Fluoro-coumarin silicon phthalocyanine, used as a sensitizer, was then encapsulated in the β-CD to decrease the electron transfer distance and thus increase the ECL signal. A competitive-type ECL immunosensor based on this system was constructed that permitted the stable and sensitive detection of deoxynivalenol. TiO2 nanomaterials can act as ECL luminophores in ECL detection methods. Deng et al. [37] reported the use of flower-like ­TiO2 nanostructures as an ECL emitter label in a PSA assay. This flower-like morphology of T ­ iO2 provides plenty of binding sites for biomolecules and nanomaterials due to its porosity and large specific surface area. In subsequent study, ­TiO2 nanoflowers@g-C3N4-Au [38] were

Fig. 4  Schematic of a dual-amplified ECL immunosensor in which TRM plays two roles. Reproduced with permission from [35] Reprinted from the journal

9

13

Topics in Current Chemistry (2020) 378:28

synthesized for an ECL immunoassay of N-terminal brain natriuretic peptide. The band levels of ­TiO2 and g-C3N4 are well matched, increasing the ECL emission efficiency. Polydopamine (PDA) linked with A ­ b2 was employed as a ECL quencher to improve detection sensitivity. Wei’s group [39] reported an ECL immunosensor for CEA based on the composite Au-FrGO-CeO2@TiO2, where Au-FrGO refers to ­Fe3O4 capped with reduced graphene oxide functionalized with Au NPs. When ­K2S2O8 was employed as a coreactant, C ­ eO2@TiO2 was found to yield greater ECL intensity than ­TiO2. The resulting ­CeO2@TiO2 heterojunction presented a smaller band gap than T ­ iO2, facilitating electron transfer between the ­CeO2@TiO2 composite and its coreactant. The Au-FrGO was included to further amplify the ECL signal. When tested, the ECL biosensor displayed a sensitive response to CEA. Tian et al. [40] presented an ECL immunosensor based on ­TiO2 nanotube arrays functionalized with graphene quantum dots (GQDs). Vertically aligned ­TiO2 nanotubes were included to provide a large surface area on which to immobilize the GQDs and antibody. The ECL intensity was enhanced sixfold after immobilizing the GQDs to form a hybrid structure (GQD/TiO2 NTs). Under optimized conditions, PSA was sensitively detected using CdTe NPs modified with ­Fe3O4 magnetic nanoparticles (CdTe/ MNPs) as a quenching label. Cui and coauthors [41] developed a potential-resolved ECL strategy for the label-free ratiometric aptasensing of cTnI that involved wrapping titanium dioxide in nanographene oxide (nGO@TiO2 NLPs) employing a “one-pot” hydrothermal method (Fig. 5a). Using K ­ 2S2O8 as a coreactant, as shown in Fig. 5b, this nanocomposite yielded dual ECL emission (ECL-1 and ECL-2) at − 1.27 V and − 1.85 V, respectively (black line). In contrast, ­TiO2 (green line) and nGO (blue line) each provided just one ECL emission peak under the same conditions. ECL-1 and ECL-2 were found to correspond with the ECL emissions of ­TiO2 and nGO by comparing ECL potentials and emission wavelengths (Fig. 5b–d), but the intensities of ECL-1 and ECL-2 were observed to be considerably stronger than the corresponding ECL emission intensities of ­TiO2 and nGO alone due to a synergistic effect. After capturing the target, the aptamer moved away from the electrode surface because of its rigidity, reducing the resistance of the electrode and enhancing the two ECL signals from the nGO@TiO2 NLPs. Furthermore, the intensity of ECL-1 and the ECL intensity of T ­ iO2 were observed to increase in an oxygen atmosphere and to decrease in a nitrogen atmosphere because some of the ­O2 was electroreduced and subsequently reacted with ­S2O82− to generate SO•− 4 . Similarly, Dai et al. [42] proposed a simple dual coreactant strategy to enhance the ECL performance of T ­ iO2 nanotubes. When K ­ 2S2O8 and ­H2O2 were added simultaneously, the ECL intensity of ­TiO2 increased 6.3-fold and 107-fold, respectively, compared to when only ­K2S2O8 or ­H2O2 was added as a coreactant, which was attributed to the increased concentration of SO•− ­ 2O2 on the electrode 4 caused by the presence of H surface. This mechanistic study of the ECL enhancement caused by dual coreactants could provide a general strategy for improving ECL-based applications of semiconductor nanomaterials. Besides acting as an immobilization substrate, T ­ iO2 catalyzes the oxidation of ­ H2O2, which makes ­ TiO2 suitable for fabricating biofunctional ECL electrodes. Wu et  al. [43] reported an enzymatic ECL choline sensor based on a ­Fe3O4-TiO2-choline oxidase (ChO) biocomposite. Choline was oxidized with

13

10

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28

Fig. 5  a Schematic of the fabrication of a label-free ratiometric ECL aptasensor. b ECL curves of the nGO@TiO2 NLPs (black line), nGO (blue line), T ­ iO2 (green line), and the FTO electrode (red line). Comparison of the ECL emission wavelengths of ECL-1/TiO2 (c) and ECL-2/nGO (d). Reproduced with permission from [41]

dissolved ­O2 by ChO to generate ­H2O2, which was then measured quantitatively based on the ECL signal from the luminol/H2O2 reaction. In this system, ­Fe3O4 nanospheres catalyzed the electrooxidation of ­H2O2. At the same time, the presence of ­TiO2 increased electron transfer and the ECL signal from luminol. Tang et  al. [44] presented a system comprising Au NPs/ionic liquid/hollowed ­TiO2 nanoshells for the sensitive ECL biosensing of cholesterol. The synthesized nanocomposite exhibited more intense luminol/H2O2 ECL emission than AuNPs or ­TiO2 alone. Cholesterol oxidase (ChOx) was immobilized on the nanofunctionalized electrode surface by glutaraldehyde (GD) and bovine serum albumin (BSA). This strategy permitted the sensitive quantification of cholesterol. A disposable biosensor [45] for glucose detection was prepared using a Au/TiO2 nanocomposite to intensify the ECL of luminol. After crosslinking glucose oxidase via GD and BSA, the ECL biosensor showed excellent stability, sensitivity, and simplicity when used for glucose detection. Moreover, Li et al. [46] reported a ECL immunosensor that used AuPdPt–MoS2@TiO2 to increase the ECL intensity of luminol by catalyzing the electrochemical reaction of ­H2O2. The nanocomposite was covalently linked directly to the amino group of luminol and the resulting

Reprinted from the journal

11

13



Topics in Current Chemistry (2020) 378:28

matrix was employed to anchor A ­ b1. Finally, insulin was sensitively measured with ­MnO2@C nanospheres used as an energy-transfer quenching label. Various applications of ECL microscopy appear in the literature due to the extremely low background light and high throughput of this technique. Because of the steric hindrance and low electrical conductivity of cells, it is difficult to perform ECL cell imaging directly. Zhu and coauthors [47] designed a direct ECL imaging strategy for single cells on a chitosan- and nano-TiO2-modified FTO glass electrode that utilized the electrocatalytic ability of nano-TiO2 (Fig. 6a). The permeable chitosan film provided a favorable microenvironment for cell immobilization and increased the space available between the cells and electrode for L012 (a luminol analog). Meanwhile, the nanosized ­TiO2 amplified the ECL signal from L012 and ­H2O2 released from the cells and permitted imaging with a high signal-to-noise ratio. Those authors subsequently investigated [48] the application of the steadystate ECL of individual rutile T ­ iO2 NPs in a biosensor targeting the local efflux from single cells (Fig.  6b). The T ­ iO2 accelerated electron transfer, enhancing the ECL intensity. In addition, oxygen vacancies in the rutile ­TiO2 adsorbed ­H2O2 and were stable from passivation by applying voltage  in the investigated range, leading to a constant ECL signal from L012/H2O2 under physiological conditions. This steadystate luminescence made it possible to visualize the H ­ 2O2 efflux from single cells using single ­TiO2 nanoparticles with high spatial and temporal resolution. Besides the luminol/H2O2 system, T ­ iO2 also strongly catalyzes other ECL systems. Yuan’s group [17] designed a system containing Ag nanoclusters/TiO2 nanoflowers (Ag NCs–TiO2 NFs) for use as a highly efficient ECL probe for the detection of amyloid-β. Due to the large surface area of the nanoflower structure and its

Fig. 6  a Schematic of the ECL imaging of cells on a chitosan- and nano-TiO2-modified electrode. b Schematic of the ECL sensing strategy involving the visualization of single T ­ iO2 NPs as a means to monitor the ­H2O2 efflux from single cells. c The ECL mechanism of Ag NCs–TiO2 NFs and comparison of the ECL intensities of the Ag NCs and the Ag NCs–TiO2 NFs. Reproduced with permission from [17, 47, 48]

13

12

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28

strongly catalytic properties, the T ­ iO2 NFs promoted the reduction of dissolved O ­2 to yield ­OH• radicals, which reacted with the Ag NCs, leading to stronger ECL emission (Fig.  6c). Ferrocene-labeled DNA was employed as an ECL quenching probe to further improve the sensitivity of the proposed biosensor. The catalysis of the reduction of dissolved ­O2 by ­TiO2 has also been applied to enhance the cathodic ECL intensity of A ­ u25 NCs in a sensitive ECL biosensor [49]. Oxygen vacancies of ­TiO2 could improve the electrocatalytic performance [50]. And using X-ray photoelectron spectroscopy, Liang et al. [51] discovered that oxygen vacancies in Cudoped ­TiO2 NPs could act as reactive centers for the conversion of dissolved ­O2 into superoxide radicals (­O2•−), which intensified the ECL of luminol. In another study, ­S2O82− was employed as a coreactant and ­TiO2 as its accelerator [52] in order to enhance the ECL efficiency of the Cu NCs used in an ultrasensitive ECL biosensor for microRNA detection. In the presence of the target, a cascade leading to signal amplification and a hybridization chain reaction were triggered, causing AT-rich double-stranded DNA to be generated on the ­TiO2. The Cu NCs that acted as the ECL luminophore in this detection system were then generated in situ on the dsDNA via A–Cu2+–T bonding. The ­TiO2 not only provided a platform for the dsDNA functionalized with Cu NCs, but it also helped to generate SO•− 4 , which improved the performance of the ECL biosensor. Zhang et al. [16] prepared a ternary ECL biosensor based on Ru(bpy)2(cpaphen)2+/TPrA/TiO2 nanoneedles for the detection of glutathione. The electrode was modified with T ­ iO2 nanoneedles in order to immobilize long dsDNA structures that adsorbed Ru(bpy)2(cpaphen)2+, the ECL luminophor. The ­TiO2 nanoneedles also acted as an accelerator for the oxidation of tripropylamine (TPrA), significantly intensifying the ECL signal from Ru(bpy)2(cpaphen)2+. When a voltage was applied to the system, electrons tunneled from the VB of ­TiO2 to the CB. The holes generated in the VB were filled by electrons from TPrA, generating T ­ PrA+•. Finally, GSH recognition was achieved with this system by including M ­ nO2 nanosheets, which were reduced by the GSH to ­Mn2+; and then Mn2+ worked as cofactor to cleave the Ru-dsDNA structures.

4 Conclusion and Future Perspectives This review has focused on the development of novel ­TiO2-based nanomaterials and innovative applications of them in PEC and ECL biosensing. Such biosensors based on ­TiO2 nanomaterials are particularly interesting mechanistically and from the perspective of biosensing applications due to the special photoelectric interconversion processes they use to probe biorecognition and biocatalytic events. However, there are still several challenges in the practical application of ­TiO2 nanomaterials to PEC and ECL biosensing. First, T ­ iO2 nanomaterials have relatively low PEC and ECL efficiencies, so other nanomaterials such as QDs or noble metal NPs are usually incorporated into the biosensing system to optimize its performance. In many cases, the strategies used to synthesize the T ­ iO2-based nanomaterials have been taken directly from energy and photocatalysis research, and the resulting nanomaterials may be not suitable for use in bioassays. More effort should be directed into designing ­TiO2 nanomaterials that are specifically for PEC and ECL biosensing. Second, Reprinted from the journal

13

13



Topics in Current Chemistry (2020) 378:28

even though there have been a few attempts to achieve bioimaging and biodetection in vivo using ­TiO2-based nanomaterials, near-infrared PEC and ECL biosensing are only rarely addressed in the literature and therefore require detailed investigation. The detection or imaging of a single molecule, cell, or particle may be an important area of research in the future. Third, in some situations, the function and mechanism of the T ­ iO2-based nanomaterial in the biosensing methods is still unclear and therefore requires further study that makes use of the advances that are currently being made in nanotechnology and instrument characterization. In a word, for the foreseeable future, the use of rationally designed and modified ­TiO2 nanomaterials in PEC and ECL biosensors has the potential to substantially broaden the range of applications of these detection systems. Acknowledgements  Funding from the National Natural Science Foundation of China (nos. 21874126 and 21675148) and The National Key Research and Development Program of China (no. 2016YFA0201300) are greatly appreciated.

References 1. Zhao W, Xu J, Chen H (2015) Photoelectrochemical bioanalysis: the state of the art. Chem Soc Rev 44(3):729–741. https​://doi.org/10.1039/C4CS0​0228H​ 2. Zang Y, Lei J, Ju H (2017) Principles and applications of photoelectrochemical sensing strategies based on biofunctionalized nanostructures. Biosens Bioelectron 96:8–16. https​://doi.org/10.1016/j. bios.2017.04.030 3. Shu J, Tang D (2019) Recent advances in photoelectrochemical sensing: from engineered photoactive materials to sensing devices and detection modes. Anal Chem. https​://doi.org/10.1021/acs.analc​ hem.9b041​99 4. Liu Z, Qi W, Xu G (2015) Recent advances in electrochemiluminescence. Chem Soc Rev 44(10):3117–3142. https​://doi.org/10.1039/C5CS0​0086F​ 5. Gao W, Saqib M, Qi L, Zhang W, Xu G (2017) Recent advances in electrochemiluminescence devices for point-of-care testing. Curr Opin Electrochem 3(1):4–10. https​://doi.org/10.1016/j.coele​ c.2017.03.003 6. Chen Y, Zhou S, Li L, Zhu J (2017) Nanomaterials-based sensitive electrochemiluminescence biosensing. Nano Today 12:98–115. https​://doi.org/10.1016/j.nanto​d.2016.12.013 7. Kumar N, Chauhan NS, Mittal A, Sharma S (2018) ­TiO2 and its composites as promising biomaterials: a review. Biometals 31(2):147–159. https​://doi.org/10.1007/s1053​4-018-0078-6 8. Zang Y, Fan J, Ju Y, Xue H, Pang H (2018) Current advances in semiconductor nanomaterialbased photoelectrochemical biosensing. Chemistry 24(53):14010–14027. https​://doi.org/10.1002/ chem.20180​1358 9. Chen J, Kong L, Sun X, Feng J, Chen Z, Fan D, Wei Q (2018) Ultrasensitive photoelectrochemical immunosensor of cardiac troponin I detection based on dual inhibition effect of Ag@Cu2O coreshell submicron-particles on CdS QDs sensitized ­TiO2 nanosheets. Biosens Bioelectron 117:340– 346. https​://doi.org/10.1016/j.bios.2018.05.037 10. Pang X, Bian H, Wang W, Liu C, Khan MS, Wang Q, Qi J, Wei Q, Du B (2017) A bio-chemical application of N-GQDs and g-C3N4 QDs sensitized T ­ iO2 nanopillars for the quantitative detection of pcDNA3-HBV. Biosens Bioelectron 91:456–464. https​://doi.org/10.1016/j.bios.2016.12.059 11. Yang L, Liu X, Li L, Zhang S, Zheng H, Tang Y, Ju H (2019) A visible light photoelectrochemical sandwich aptasensor for adenosine triphosphate based on ­MgIn2S4–TiO2 nanoarray heterojunction. Biosens Bioelectron 142:111487. https​://doi.org/10.1016/j.bios.2019.11148​7 12. Shu J, Qiu Z, Zhuang J, Xu M, Tang D (2015) In situ generation of electron donor to assist signal amplification on porphyrin-sensitized titanium dioxide nanostructures for ultrasensitive photoelectrochemical immunoassay. ACS Appl Mater Interfaces 7(42):23812–23818. https​://doi.org/10.1021/ acsam​i.5b087​42

13

14

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28



13. Liu X, Huo X, Liu P, Tang Y, Xu J, Liu X, Zhou Y (2017) Assembly of M ­ oS2 nanosheet-TiO2 nanorod heterostructure as sensor scaffold for photoelectrochemical biosensing. Electrochim Acta 242:327–336. https​://doi.org/10.1016/j.elect​acta.2017.05.037 14. Da P, Li W, Lin X, Wang Y, Tang J, Zheng G (2014) Surface plasmon resonance enhanced real-time photoelectrochemical protein sensing by gold nanoparticle-decorated ­TiO2 nanowires. Anal Chem 86(13):6633–6639. https​://doi.org/10.1021/ac501​406x 15. Pang X, Bian H, Su M, Ren Y, Qi J, Ma H, Wu D, Hu L, Du B, Wei Q (2017) Photoelectrochemical cytosensing of RAW264.7 macrophage cells based on a ­TiO2 nanoneedls@MoO3 array. Anal Chem 89(15):7950–7957. https​://doi.org/10.1021/acs.analc​hem.7b010​38 16. Zhang R, Zhong X, Chen A, Liu J, Li S, Chai Y, Zhuo Y, Yuan R (2019) Novel Ru(bpy)2(cpaphen)2+/ TPrA/TiO2 ternary ECL system: an efficient platform for the detection of glutathione with M ­ n2+ as substitute target. Anal Chem 91(5):3681–3686. https​://doi.org/10.1021/acs.analc​hem.8b057​95 17. Zhou Y, Wang H, Zhuo Y, Chai Y, Yuan R (2017) Highly efficient electrochemiluminescent silver nanoclusters/titanium oxide nanomaterials as a signal probe for ferrocene-driven light switch bioanalysis. Anal Chem 89(6):3732–3738. https​://doi.org/10.1021/acs.analc​hem.7b000​90 18. Tang Y, Liu X, Zheng H, Yang L, Li L, Zhang S, Zhou Y, Alwarappan S (2019) A photoelectrochemical aptasensor for aflatoxin B1 detection based on an energy transfer strategy between CeTiO2@MoSe2 and Au nanoparticles. Nanoscale 11(18):9115–9124. https​://doi.org/10.1039/C9NR0​ 1960J​ 19. Liu N, Chen S, Li Y, Dai H, Lin Y (2017) Self-enhanced photocathodic matrix based on poly-dopamine sensitized T ­ iO2 mesocrystals for mycotoxin detection assisted by a dual amplificatory nanotag. New J Chem 41(9):3380–3386. https​://doi.org/10.1039/C6NJ0​3157A​ 20. Dai H, Xu G, Zhang S, Hong Z, Lin Y (2015) A ratiometric biosensor for metallothionein based on a dual heterogeneous electro-chemiluminescent response from a ­TiO2 mesocrystalline interface. Chem Commun 51(36):7697–7700. https​://doi.org/10.1039/C5CC0​1402F​ 21. Liu P, Liu X, Huo X, Tang Y, Xu J, Ju H (2017) T ­ iO2–BiVO4 heterostructure to enhance photoelectrochemical efficiency for sensitive aptasensing. ACS Appl Mater Interfaces 9(32):27185–27192. https​://doi.org/10.1021/acsam​i.7b070​47 22. Fu B, Wu W, Gan L, Zhang Z (2019) Bulk/surface defects engineered ­TiO2 nanotube photonic crystals coupled with plasmonic gold nanoparticles for effective in vivo near-infrared light photoelectrochemical detection. Anal Chem 91(22):14611–14617. https​://doi.org/10.1021/acs.analc​hem.9b037​ 33 23. Zhao W, Wang J, Zhu Y, Xu J, Chen H (2015) Quantum dots: electrochemiluminescent and photoelectrochemical bioanalysis. Anal Chem 87(19):9520–9531. https​://doi.org/10.1021/acs.analc​ hem.5b004​97 24. Cai G, Yu Z, Ren R, Tang D (2018) Exciton-plasmon interaction between AuNPs/graphene nanohybrids and CdS quantum dots/TiO2 for photoelectrochemical aptasensing of prostate-specific antigen. ACS Sensors 3(3):632–639. https​://doi.org/10.1021/acsse​nsors​.7b008​99 25. Wang L, Meng Y, Zhang C, Xiao H, Li Y, Tan Y, Xie Q (2019) Improving photovoltaic and enzymatic sensing performance by coupling a core-shell Au nanorod@TiO2 heterostructure with the bioinspired l-DOPA polymer. ACS Appl Mater Interfaces 11(9):9394–9404. https​://doi. org/10.1021/acsam​i.8b192​84 26. Yan Z, Wang Z, Miao Z, Liu Y (2016) Dye-sensitized and localized surface plasmon resonance enhanced visible-light photoelectrochemical biosensors for highly sensitive analysis of protein kinase activity. Anal Chem 88(1):922–929. https​://doi.org/10.1021/acs.analc​hem.5b036​61 27. Halawa MI, Lai J, Xu G (2018) Gold nanoclusters: synthetic strategies and recent advances in fluorescent sensing. Materials Today Nano 3:9–27 28. Zhu J, Huo X, Liu X, Ju H (2016) Gold nanoparticles deposited polyaniline–TiO2 nanotube for surface plasmon resonance enhanced photoelectrochemical biosensing. ACS Appl Mater Interfaces 8(1):341–349. https​://doi.org/10.1021/acsam​i.5b088​37 29. Hao N, Hua R, Chen S, Zhang Y, Zhou Z, Qian J, Liu Q, Wang K (2018) Multiple signal-amplification via Ag and ­TiO2 decorated 3D nitrogen doped graphene hydrogel for fabricating sensitive label-free photoelectrochemical thrombin aptasensor. Biosens Bioelectron 101:14–20. https​://doi. org/10.1016/j.bios.2017.10.014 30. Wang Y, Zhao G, Zhang Y, Du B, Wei Q (2019) Ultrasensitive photoelectrochemical immunosensor based on Cu-doped T ­ iO2 and carbon nitride for detection of carcinoembryonic antigen. Carbon 146:276–283. https​://doi.org/10.1016/j.carbo​n.2019.02.008

Reprinted from the journal

15

13



Topics in Current Chemistry (2020) 378:28

31. Zhang Y, Li M, Wang H, Yuan R, Wei S (2019) Supersensitive photoelectrochemical aptasensor based on Br, N-codoped T ­ iO2 sensitized by quantum dots. Anal Chem 91(16):10864–10869. https​://doi.org/10.1021/acs.analc​hem.9b026​00 32. Shu J, Qiu Z, Lv S, Zhang K, Tang D (2018) Plasmonic enhancement coupling with defect-engineered ­TiO2–x: a mode for sensitive photoelectrochemical biosensing. Anal Chem 90(4):2425– 2429. https​://doi.org/10.1021/acs.analc​hem.7b052​96 33. Qiu Z, Shu J, Tang D (2018) Near-infrared-to-ultraviolet light-mediated photoelectrochemical aptasensing platform for cancer biomarker based on core–shell NaYF4:Yb, Tm@TiO2 upconversion microrods. Anal Chem 90(1):1021–1028. https​://doi.org/10.1021/acs.analc​hem.7b044​79 34. Tsuneyasu S, Ichihara K, Nakamura K, Kobayashi N (2016) Why were alternating-currentdriven electrochemiluminescence properties from Ru(bpy)32+ dramatically improved by the addition of titanium dioxide nanoparticles? Phys Chem Chem Phys 18(24):16317–16324. https​://doi. org/10.1039/C6CP0​2881K​ 35. Zheng H, Yi H, Lin W, Dai H, Hong Z, Lin Y, Li X (2018) A dual-amplified electrochemiluminescence immunosensor constructed on dual-roles of rutile ­TiO2 mesocrystals for ultrasensitive zearalenone detection. Electrochim Acta 260:847–854. https​://doi.org/10.1016/j.elect​ acta.2017.12.054 36. Zheng H, Yi H, Dai H, Fang D, Hong Z, Lin D, Zheng X, Lin Y (2018) Fluoro-coumarin silicon phthalocyanine sensitized integrated electrochemiluminescence bioprobe constructed on ­TiO2 MOFs for the sensing of deoxynivalenol. Sens Actuators B Chem 269:27–35. https​://doi. org/10.1016/j.snb.2018.04.149 37. Deng W, Chu C, Ge S, Yu J, Yan M, Song X (2015) Electrochemiluminescence PSA assay using an ITO electrode modified with gold and palladium, and flower-like titanium dioxide microparticles as ECL labels. Microchim Acta 182(5):1009–1016. https​://doi.org/10.1007/s0060​ 4-014-1423-2 38. Zhao Y, Li L, Hu L, Zhang Y, Wu D, Ma H, Wei Q (2019) An electrochemiluminescence immunosensor for the N-terminal brain natriuretic peptide based on the high quenching ability of polydopamine. Microchim Acta 186(9):606. https​://doi.org/10.1007/s0060​4-019-3709-x 39. Yang L, Zhu W, Ren X, Khan MS, Zhang Y, Du B, Wei Q (2017) Macroporous graphene capped ­Fe3O4 for amplified electrochemiluminescence immunosensing of carcinoembryonic antigen detection based on ­ CeO2@TiO2. Biosens Bioelectron 91:842–848. https​://doi.org/10.1016/j. bios.2017.01.055 40. Tian C, Wang L, Luan F, Zhuang X (2019) An electrochemiluminescence sensor for the detection of prostate protein antigen based on the graphene quantum dots infilled ­TiO2 nanotube arrays. Talanta 191:103–108. https​://doi.org/10.1016/j.talan​ta.2018.08.050 41. Han Z, Shu J, Liang X, Cui H (2019) Label-free ratiometric electrochemiluminescence aptasensor based on nanographene oxide wrapped titanium dioxide nanoparticles with potential-resolved electrochemiluminescence. Anal Chem 91(19):12260–12267. https​://doi.org/10.1021/acs.analc​ hem.9b023​18 42. Dai P, Yu T, Shi H, Xu J, Chen H (2015) General strategy for enhancing electrochemiluminescence of semiconductor nanocrystals by hydrogen peroxide and potassium persulfate as dual coreactants. Anal Chem 87(24):12372–12379. https​://doi.org/10.1021/acs.analc​hem.5b038​90 43. Wu X, Chai Y, Yuan R, Liang W, Yuan D (2014) A novel electrochemiluminescence choline biosensor based on biofunctional AMs-ChO biocomposite. Sens Actuators B Chem 204:429–436. https​ ://doi.org/10.1016/j.snb.2014.07.125 44. Tang S, Zhao Q, Tu Y (2016) A sensitive electrochemiluminescent cholesterol biosensor based on Au/hollowed-TiO2 nano-composite pre-functionalized electrode. Sens Actuators B Chem 237:416– 422. https​://doi.org/10.1016/j.snb.2016.06.110 45. Yu L, Wei X, Fang C, Tu Y (2016) A disposable biosensor for noninvasive diabetic diagnosis rest on the Au/TiO2 nano-composite intensified electrochemiluminescence. Electrochim Acta 211:27– 35. https​://doi.org/10.1016/j.elect​acta.2016.06.034 46. Li X, Sun X, Fan D, Yan T, Feng R, Wang H, Wu D, Wei Q (2019) A ternary quenching electrochemiluminescence insulin immunosensor based on M ­ n2+ released from M ­ nO2@carbon core-shell nanospheres with ascorbic acid quenching AuPdPt–MoS2@TiO2 enhanced luminol. Biosens Bioelectron 142:111551. https​://doi.org/10.1016/j.bios.2019.11155​1 47. Liu G, Ma C, Jin B, Chen Z, Zhu J (2018) Direct electrochemiluminescence imaging of a single cell on a chitosan film modified electrode. Anal Chem 90(7):4801–4806. https​://doi.org/10.1021/ acs.analc​hem.8b001​94

13

16

Reprinted from the journal

Topics in Current Chemistry (2020) 378:28 48. Cui C, Chen Y, Jiang D, Chen H, Zhang J, Zhu J (2019) Steady-state electrochemiluminescence at single semiconductive titanium dioxide nanoparticles for local sensing of single cells. Anal Chem 91(1):1121–1125. https​://doi.org/10.1021/acs.analc​hem.8b047​78 49. Zhou Y, Chai Y, Yuan R (2019) Highly efficient dual-polar electrochemiluminescence from ­Au25 nanoclusters: the next generation of multibiomarker detection in a single step. Anal Chem 91(22):14618–14623. https​://doi.org/10.1021/acs.analc​hem.9b037​36 50. Zeng X, Bai Y, Choi SM, Tong L, Aleisa RM, Li Z, Liu X, Yu R, Myung NV, Yin Y (2019) Mesoporous TiO2 nanospheres loaded with highly dispersed Pd nanoparticles for pH-universal hydrogen evolution reaction. Materials Today Nano 6:100038 51. Liang J, Xu Q, Teng X, Guan W, Lu C (2019) Superoxide-triggered luminol electrochemiluminescence for detection of oxygen vacancy in oxides. Anal Chem. https​://doi.org/10.1021/acs.analc​ hem.9b051​56 52. Liao H, Zhou Y, Chai Y, Yuan R (2018) An ultrasensitive electrochemiluminescence biosensor for detection of microRNA by in situ electrochemically generated copper nanoclusters as luminophore and ­TiO2 as coreaction accelerator. Biosens Bioelectron 114:10–14. https​://doi.org/10.1016/j. bios.2018.05.011

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Reprinted from the journal

17

13

Topics in Current Chemistry (2020) 378:13 https://doi.org/10.1007/s41061-019-0277-9 REVIEW

DNA–Iron Oxide Nanoparticles Conjugates: Functional Magnetic Nanoplatforms in Biomedical Applications José Raúl Sosa‑Acosta1,2 · Claudia Iriarte‑Mesa1 · Greter A. Ortega1,3 · Alicia M. Díaz‑García1  Received: 24 July 2019 / Accepted: 20 December 2019 / Published online: 10 January 2020 © Springer Nature Switzerland AG 2020

Abstract The use of magnetic nanoparticles (MNPs), such as iron oxide nanoparticles (IONPs), in biomedicine is considered to be a valuable alternative to the more traditional materials due to their chemical stability, cost-effectiveness, surface functionalization, and the possibility to selectively attach and transport targeted species to the desired location under a magnetic field. One of the many main applications of MNPs is DNA separation, which enables genetic material manipulation; consequently, MNPs are used in numerous biotechnological methods, such as gene transfection and molecular recognition systems. In addition, the interaction between the surfaces of MNPs and DNA molecules and the magnetic nature of the resulting composite have facilitated the development of safe and effective gene delivery vectors to treat significant diseases, such as cancer and neurological disorders. Furthermore, the special recognition properties of nucleic acids based on the binding capacity of DNA and the magnetic behavior of the nanoparticles allowing magnetic separation and concentration of analytes have led to the development of biosensors and diagnostic assays; however, both of these applications face important challenges in terms of the improvement of selective nanocarriers and biosensing capacity. In this review, we discuss some aspects of the properties and surface functionalization of MNPs, the interactions between DNA and IONPs, the preparation of DNA nanoplatforms and their biotechnological applications, such as the magnetic separation of DNA, magnetofection, preparation of DNA vaccines, and molecular recognition tools. Keywords  Magnetic nanoparticles · DNA conjugation · Nucleic acid separation · DNA-based therapeutics

Chapter 2 was originally published as Sosa‑Acosta, J. R., Iriarte‑Mesa, C., Ortega, G. A. & Díaz‑García, A. M. Topics in Current Chemistry (2020) 378: 13. https://doi.org/10.1007/s41061-019-0277-9. * Alicia M. Díaz‑García [email protected] Extended author information available on the last page of the article Reprinted from the journal

19

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:13

Abbreviations APTES (3-Aminopropyl) triethoxysilane CuAAC​ Cu(I)-catalyzed azide-alkyne cycloaddition IONPs Iron oxide nanoparticles MNPs Magnetic nanoparticles ODN Oligonucleotide PAMAM Polyamidoamine dendrimers PEG Polyethylene glycol PEI Polyethylenimine pHEMA Poly(2-hydroxyethyl methacrylate) PNA 4-Pyridyldithiol-derivatized peptide nucleic acid siRNAs Small interfering RNAs ssDNA Single-stranded DNA ssODN Single-stranded oligonucleotide TEOS Tetraethyl orthosilane TMOS Tetramethyl orthosilane

1 Introduction The increasing amount of scientific research that is focusing on nanomaterials has resulted in significant progress in many practical applications. This class of compounds is referred to as ‘nanoscaled’ due to the size of the particles, a term which means that at least one of the three dimensions is in the range of 1–100  nm. As a multipurpose science, nanotechnology has extended the interest of researchers into novel systems at this small scale due to the exceptional properties and applications of the nanomaterials. Nanomaterials comprise many materials, such as carbon nanotubes, fullerenes, nanocomposites, nanopolymers, nanovectors, nanoparticles, nanofibers, nanowires, nanorods, among many others [1]. The primary applications of these materials are in the technological and biomedical fields, including water treatment [2, 3], catalysis and electrocatalysis [4–10], air purification [11], photovoltaics [12], cancer treatment [13], among others. Magnetic nanoparticles (MNPs) are considered to be the center of nanotechnology-based structures and have had a substantial impact in the fields of nanomedicine, analytical chemistry, electronics, and biosensing [14–17]. To date, significant improvements have been made in the synthesis and characterization of such systems, with a focus on achieving and maintaining a desired size, morphology, composition, and surface chemistry. The use of MNPs and in particular ferrite colloids in the field of biomedicine is associated with their physical properties, magnetic susceptibility, biocompatibility, and low toxicity [18]. The surface functionalization of such materials allows many structures to be designed while taking into consideration the conjugated (bio)molecule and the specific target. The conjugation of MNPs to DNA fragments is just one example of unique magnetic properties and biological selectivity combinations that are aimed at improving the efficiency of diagnosis and therapy of diseases [19, 20]. Several approaches to conjugate nucleic acids with MNPs have been reported

13

20

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13

in the literature [21–23]; these include the direct binding of a nucleic acid molecule to MNPs or the formation of chemical bonds that requires preliminary modification of the nanoparticles’ surface and/or DNA molecules. Thus, the design of DNA-based magnetic nanocomposites for applications in nanomedicine is not easily implemented and depends on the ultimate purpose for which that nanocomposite is intended [24–27]. Over the last decades, a number of authors have pointed out that magnetofection or gene delivery is a fundamental stage in nanomedicine development [28–30]. It is well known that one of the fundamental steps of gene delivery is specificity in DNA separation. However, such methodology is also important in many other applications, such as magnetosensitive biosensors [31], theranostics [32], and vaccine preparation [33]. Even though the DNA isolation process itself remains a challenge in terms of optimization and solid phase support selection, the use of magnetic separation is still advantageous compared to traditional techniques [34]. The aim of this review article is to briefly outline the main properties of MNPs and MNPs/DNA biocomposites. Current information on synthesis methods, surface modification, and DNA interactions are also discussed. Additionally, we consider a systematic description of the DNA isolation process using MNPs and the functioning principles of other nucleic acid-based nanobiohybrid systems.

2 Synthesis, Properties, and Surface Functionalization Strategies of Magnetic Nanoparticles The term magnetic nanoparticle covers a wide spectrum of nanostructured materials that have the advantageous property of being magnetic in nature, thereby enabling their use for different applications. This description covers a wide range of nanoparticles, including metallic [35], bimetallic [36], and metal oxide nanoparticles [37], in diverse architectures, such as core–shell structures [38] or Janus-type nanoparticles [39]. Among these structures are the iron oxide systems that have been intensively studied for biomedical and technological applications [29]. Magnetic iron oxide nanoparticles (IONPs) have significant advantages because they are inexpensive to produce, exhibit sufficient physical and chemical stability, and have sufficient biocompatibility [15]. These properties together with a proper magnetic response are the factors determining the use of IONPs in targeted drug delivery, hyperthermia, magnetic resonance imaging, detection of cancer biomarkers, clinical diagnosis, bioremediation, and DNA isolation [40–43], among others. (Fig. 1). Particle size control, phase purity, colloidal stability, and magnetic nature have been the focus of attention during the development of methodologies [44–46]. These features are fundamental to achieving an appropriate nanocolloid and therefore fulfilling the requirements for its use in practical applications. The main pathways for the synthesis of IONPs, such as magnetite (­ Fe3O4), can be classified as: (1) physical methods, such as gas-phase deposition and electron beam lithography, which are difficult techniques in terms of controlling particle size [47]; (2) chemical preparation methods, such as sol–gel, thermal decomposition, chemical coprecipitation, hydrothermal reactions, flow injection, electrochemical, and syntheses using nanoreactors Reprinted from the journal

21

13

Topics in Current Chemistry (2020) 378:13

Fig. 1  Schematic illustration of the main applications of magnetic nanoparticles (MNPs) as a function of their cargo or carrier characteristics

[48, 49]; and (3) microbial methods, which ensure good reproducibility and scalability at a low cost and moderate preparation temperatures [50, 51]. Among the chemical routes, coprecipitation is considered to be the simplest, cheapest and most environmentally friendly procedure. This route involves the simultaneous precipitation of ­Fe2+ and ­Fe3+ ions in basic aqueous media. The reaction temperature is limited by the boiling point of water, and the IONPs synthesized under these conditions usually exhibit a low degree of crystallinity and large polydispersity [52, 53]. Because of the large number of possible synthesis parameters, the coprecipitation method makes it possible to study how the final properties of IONPs can be controlled by various factors, such as the nature of the starting reagents, molar ratio of ­Fe2+ to ­Fe3+, alkali type, pH, stirring, ionic strength, surfactant, and temperature [54]. For example, the slow addition of an ammonia solution typically results in an increased size of nanoparticles, whereas the fast addition leads to slightly smaller ones. Compared with coprecipitation, thermal decomposition is a useful technique by which to prepare colloidally stable nanoparticles with a narrow particle size distribution (Fig.  2). In this methodology, the reaction mixture typically consists of an organometallic precursor as a metal source, surfactants, and an organic solvent with a high boiling point [55]. Airless synthetic techniques are often required when this pathway is employed  due to the use of air-sensitive molecular precursors, and the method cannot be regarded as the most environmentally friendly one due to the use of often toxic chemicals during synthesis [55, 56]. However, the high temperatures at which the reactions take place and the presence of amphiphilic surfactant molecules in a non-polar medium limit the use of nanoparticles produced in this way in most biomedical applications unless modification occurs. The hydrothermal or solvothermal technique is also considered to be employ high-temperature pathways to obtain magnetic nanocolloids. This method is dominated by the classical synthesis of nanoparticles via coprecipitation, followed by the growth of the particles under hydrothermal conditions, which ensures a high crystallinity degree as well as magnetization values. However, the use of a sealed Teflon container and the heating treatment above the boiling temperature of the water allow the production of IONPs with a broader particle size distribution in comparison to thermal decomposition products [54, 56].

13

22

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13

Fig. 2  Iron oxide nanoparticles (IONPs) coated with oleic acid synthesized by coprecipitation (a) and thermal decomposition (b) (authors’ own unpublished results)

IONPs are not the only iron-containing materials with important magnetic properties. Spinel ferrites are, in general, complex systems derived from iron oxides that are chemically combined with one or more metallic elements to improve their magnetic response. Such nanostructures have a common component and formula, ­MFe2O4, where M can be C ­ o2+, ­Mn2+, ­Ni2+, ­Zn2+, and others divalent metal ions [15, 57]. Metal-doped iron oxides with a spinel structure have been prepared by a high-temperature reaction between corresponding divalent metal chloride and iron tris-2,4-pentadioate [58], but various other synthetic methods have been reported, such as electrospinning [57], coprecipitation [59], and polyol [15]. In addition to preventing agglomeration and enhance colloidal stability, the functionalization of IONPs allows higher water compatibility and better magnetic control. Surface modification is also necessary to the conjugation of biologically active substances, an important process for nanomaterial applications to biology and medicine. Chemical modification methods for the surface of MNPs can be conditionally divided into two groups, noncovalent (or nonspecific) and covalent modification [60]. Such processes encompass ligand addition [61], ligand exchange [62], and encapsulation of diverse materials, including small molecules [63], organic or polymeric ligands [64, 65], dense polymer matrix [66], and inorganic materials [67]. In general, so-called core–shell structures representing a magnetic core grafted with a layer of a polymer or inorganic material are the most widely used MNPs. They combine the properties of both materials in a single smart and multifunctional system [68, 69]. The coating method is dependent on the nature of the grafting materials and the intended final application. Surface modification of IONPs is normally achieved in a single step during the synthesis procedure or subsequently in a post-synthesis protocol. In the former case, the coating process starts as soon as nucleation occurs, preventing further particle growth. Several ligands have been used for direct functionalization, including carboxylates [70], phosphonates [71], thiol [72], or amino groups [52]. Molecules such as citrate and amino acids may Reprinted from the journal

23

13

Topics in Current Chemistry (2020) 378:13

be adsorbed onto the surface of the IONPs via coordination of carboxylate functionalities, thereby exposing new terminal groups for further modification [73]. The post-synthesis protocols are always divided into two steps: the first step consists of IONPs synthesis, followed by surface modification as a separate step. Such techniques are performed mainly via ligand addition or exchange and encapsulation using polymeric matrixes [74]. One advantage of these methodologies is the possibility to design multiple nanostructures using either one protocol or another as well as a combination of both. Some examples of magnetic platforms based on IONPs for the conjugation of DNA are given in Table 1. The synthesis and coating methods are also presented, showing the diversity of strategies used for obtaining functional conjugates.

3 Interaction of DNA–Magnetic Nanoparticles: Preparation of Nanoplatforms DNA molecules play an important role in transferring genetic information through generations. Such structures consist of several nucleosides sharing a phosphate backbone with sequences generated from assemblies of bases able to form a double helix structure [84]. Due to its programmability, cost-effectiveness, ease of modification, and the ability to recognize a broad range of analytes, DNA is a highly attractive molecule for use in designing hybrid materials [85]. Such features together with the extraordinary properties of MNPs allow the fabrication of nanoplatforms as powerful molecular recognition tools and targeted drug delivery carriers [86, 87]. Nanoplatform design depends on the type of interaction between DNA and the nanoparticle surface. Since a large number of coupling agents are commercially available, the covalent immobilization of nucleic acids, as well as of other biomolecules, is easy to achieve (Fig. 3). Such methodologies include traditional methods of bioconjugation and the ‘click’-chemistry approaches, such as carbodiimide activation, thiol-disulfide exchange, aldehyde-amine condensation, and azide-alkyne cycloaddition [88, 89]. Covalent immobilization is also attained with thiolated and aminated molecules [90]. To this end, amino, sulfhydryl, carboxyl, and azido groups are initially formed on the surface of the nanoparticles. Although all of these methodologies are well described, some key parameters need to be considered, such as biomolecule orientation and the specific activity [17]. The covalent immobilization of DNA onto the surface of MNPs is not the only interaction of such biocomposites, but it is certainly considered to be a very useful alternative for in  vitro diagnosis [24]. However, different methods are often employed to achieve bioconjugation, such as physical adsorption, Van der Waals, electrostatic or high-affinity noncovalent interactions [25–27]. Knowledge of the relevant adsorption mechanisms provides valuable information in terms of nanomaterial surface design with the aim to manipulate or suppress highly specific binding and to control bonds between the components of the biocomposites [17]. For example, the use of an electrostatic approach could have advantages in terms of time and resources, but a high-affinity methodology could be more specific.

13

24

Reprinted from the journal

Reprinted from the journal

25

Specific adsorption of DNA

Molecular recognition

DNA exaction for prenatal diagnosis

Glutaraldehyde cross-linking Immobilization of enzymes

Electrostatic adsorption

Detection of cancer biomarkers

Gene delivery

DNA extraction

Gene detection

DNA storage

Detection of mutated DNA

DNA exaction

Application

[83]

[82]

[81]

[80]

[79]

[67]

[78]

[77]

[76]

[75]

References

APTES, 3-(Aminopropyl) triethoxysilane; PEI, polyethyleneimine; ssDNA, single-stranded DNA; CaP, calcium phosphate; PEG, polyethylene glycol; PAA, polyacrylic acid; PDA, polydopamine

5-Methyl cytosine antibody (Abcam, Cambridge, UK)

PDA

Hydrothermal

Co-precipitation

Fe3O4

Core–shell gold-coated magnetic nanoparticles ­(Fe3O4-Au-NP)

Co-precipitation PAA and free radical polymerization

Fe3O4

Encapsulation (siRNA)

Adsorption

Covalent coupling

Encapsulation

Molecular recognition

Electrostatic adsorption

DNA interaction

PAA–poly(PEG) block-copolymer Covalent coupling (PAA-b-PEGMA)

CaP and PEG-polyanion block copolymers

Co-precipitation

Co-precipitation

Fe3O4 coated with silica shell

Fe3O4

Silica

Labeled ssDNA

Solvothermal

Fe3O4@SiO2@Au

Solvothermal

PEI



Fe3O4@SiO2

Thiolated DNA

Co-precipitation

APTES

Co-precipitation and reverse microemulsion

Synthetic method Coating agent

Graphene-coated iron nanoparticles (Fe/C) with a sulfonate functionalization

Au-Fe3O4 yolk-shell

Fe3O4

Magnetic nanoparticles

Table 1  Magnetic platforms based on DNA-conjugated iron oxide nanoparticles

Topics in Current Chemistry (2020) 378:13  

13



Topics in Current Chemistry (2020) 378:13

Fig. 3  Some examples of bioconjugation using the covalent approach. Left, the use of coupling agents, such as carbodiimide [90] and ­SOCl2 [91] is shown; right, the direct bioconjugation approach using biomolecule-modified acetic anhydride- (A) [17], N-hydroxysuccinimide- (B) [88], and epoxy group- (C) functionalized biomolecules [17] is shown. NHS N-Hydroxysuccinimide

Electrostatic or the Coulomb interaction between IONPs and negatively charged DNA is the main functioning principle associated with DNA isolation. Modification with cationic materials is fundamental to the preparation of such systems. The use of positively charge nanoparticles in DNA immobilization seems to be a very common alternative in non-viral gene delivery systems due to the effectiveness and ease for automation of such nanoparticles [92]. In this regard, many different materials, such as cationic lipids or ionic liquids, proteins, and polycations, have been intensively studied [27, 29]. For example, Pandit et al. reported the functionalization of iron oxide microparticles with chitosan polymer for DNA purification [93]. In their article, the authors reported the capture of DNA at a pH optimal for PCR, enabling direct amplification from the microparticles. Classical adsorbents, such as silica and alkoxysilanes, have also been studied for use in DNA isolation [94, 95]. The advantages of these adsorbents are related to their biocompatibility, chemical stability, and nucleic acid affinity. Taking into consideration the same electrostatic principle, aminosilane-coated MNPs interact with the polyanionic DNA molecule due to the presence of phosphate groups in DNA and positive functional groups on the surface of nanoparticles. Generally, a silica coating is achieved through the sol–gel reaction (also known as the Stöber process), in which silica is synthesized via the hydrolysis and condensation of silicon orthoester [Si(OR)4], such as tetraethyl orthosilane (TEOS) and tetramethyl orthosilane (TMOS)] [50, 96]. A summary of the most important interactions of DNA with the surface of MNPs is given in Fig. 4. Since nitrogen-containing bases in DNA can strongly coordinate to the surface of ­Fe3O4 [97], the preparation of nanoparticles by coprecipitation onto DNA molecules is also considered to be a bioconjugation methodology. This procedure allows the design of one-dimensional (1D) nanostructured architectures due to the templating features of DNA. In this context, Byrne et al. studied the coprecipitation

13

26

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13



Fig. 4  DNA interactions with MNPs. a, b An electrostatic approach using unmodified (a) and functionalized (b) nanoparticles. c Coprecipitation of IONPs on DNA. d Formation of complementary hydrogen bonds between oligonucleotides covalently immobilized on MNPs and target DNA. e Noncovalent highspecific interactions

of nanoparticles on DNA and showed that denatured herring sperm DNA acts as a template for the preparation of magnetic nanowires [98]. Similarly, Hasan et  al. reported the use of the coprecipitation approach in the preparation of magnetic and conductive nanowires by DNA templating [99]. In this latter case, DNA proved to be a highly effective template for controlling the metal oxide formation, confining its growth in two dimensions to yield structurally well-defined, high-aspect-ratio nanowires with diameters of up to 30  nm. Such nanoplatforms could be useful in nanoelectronics, photonics, chemical sensors, and biological probes. The formation of hydrogen bonds between complementary DNA strands is a basic principle in the fabrication of nucleic acid–MNP hybrid nanocomposites by self-assembly. The complementary binding of an oligonucleotide (ODN) immobilized on MNPs and the target nucleic acid molecule (known as the hybridization process) underlies the operating principles of some diagnostic assays and biosensors (Fig.  5). However, the critical step during the preparation of ODN-modified MNPs is the conjugation to the nanoparticle surface. The use of covalent bonds is

Fig. 5  Schematic procedure of the hybridization process in nucleic acid–MNP hybrid nanocomposites and their applications in specific DNA sensing. ODN Oligonucleotide, QDs quantum dots Reprinted from the journal

27

13



Topics in Current Chemistry (2020) 378:13

preferred to electrostatic interaction due to the high specific binding. Taking this fact into account, the manipulation of click-reactions, phosphoramidate bond formation, thiolated DNA, and 4-pyridyldithiol-derivatized peptide nucleic acid (PNA) are taken into consideration in the fabrication of DNA hybridization biosensors [23, 100–102]. According to Robinson et  al., many possibilities are derived by exploiting the chemistry of gold–sulfur (Au–S) complexes [23]. These authors reported the synthesis and characterization of Co–Au and ­Fe3O4@Au core–shell nanoparticles by reducing a gold (III) salt in a dispersion of such MNPs. The presence of the Au shell leads the functionalization of nanoparticles with thiolated single-stranded DNA (ssDNA) and subsequently its use for hybridization processes. The Cu(I)-catalyzed azide-alkyne cycloaddition (CuAAC) “click” reaction has also been employed to achieve the attachment of DNA to ferrite nanoparticles. For example, the research group of Sreenivasulu have studied the self-assembly of multiferroic nanocomposites using DNA–DNA hybridization [100]. In this case, the azide-alkyne cycloaddition mediates the interaction between azide-functionalized barium titanate or nickel ferrite nanoparticles and alkyne-modified single-stranded oligonucleotide (ssODN). Such complex hybridized nanocomposites allow the combination of ferroelectric and ferromagnetic phases to study the mechanical strain of the related system. Along with the above-mentioned approaches, the covalent conjugation of DNA fragments to MNPs by forming a carbodiimide-mediated phosphoramidate bond is considered to be another useful route for developing hybridization biosensors. For example, Zhu et al. have investigated the application of such nanocomposites for electrochemical DNA hybridization [101], taking advantage of ssDNA immobilization on MNPs and using zinc sulfide nanoparticles as the oligonucleotide label. In most of the above-mentioned reports, oligonucleotides were used as conventional sequence-specific fragments. However, although Watson–Crick base pairing is remarkably specific, the mismatch discrimination of the ODN recognizer is not sufficiently selective and, in addition, it is susceptible to hydrolyzation by endogenous nucleases and proteases. To overcome such limitations during biomedical applications, the strategy of utilizing 4-pyridyldithiol-derivatized PNA as the sequence-specific gene recognizer can be considered. As a DNA analog, PNA comprises a polyamine instead of a sugar-phosphate backbone; as such, it is capable of binding DNA oligomers following Watson–Crick base pairing rules to form a PNA–DNA duplex that is significantly more stable than the corresponding DNA–DNA duplex. Wang et al. studied the use of PNA–MNPs biocomposites in gene recognition using surface-enhanced Raman scattering [102] and reported that PNA-modified MNPs can be easily prepared via a thiol-disulfide exchange reaction followed by the hybridization protocol. While Coulombic interactions of DNA with charged surfaces show a lack of specificity, the avidin–biotin complex is considered to be one of the most specific and stable noncovalent interactions. Since conjugations of biotin and its binding proteins (avidin and analogs such as streptavidin and neutravidin) are stable and highly specific and do not involve unstable intermediates, they could serve as a promising tool in biomedical and nanotechnological applications. Avidin is a basic tetrameric glycoprotein composed of four identical subunits that binds to biotin with

13

28

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13



high specificity and affinity (about 1­ 03- to 1­ 06-fold higher than an antigen–antibody interaction [103]). Biotin-based conjugates are also easy to synthesize and have less impact on the activity of the biomolecules. However, despite its enormous advantages and wide applicability, avidin has several limitations, including non-specific binding and possible immunogenicity. To avoid these limitations, concentrated efforts have been devoted to discovering and engineering superior variants of avidin. Information on the structure and physical–chemical properties of such variants are described in detail by de Freitas et al. [25, 104]. The specific binding of MNPs to DNA using high-affinity biotin–streptavidin noncovalent interactions requires preliminary modification of the surface of MNPs and nucleic acids with the appropriate molecules [17]. For example, Cannon et al. designed streptavidin-modified IONPs to bind biotinylated ssDNA-labeled with a Cy3 fluorescent dye [105] and thus enhance the immobilization process. The authors concluded that such nanocomposites are sufficiently specific to allow amplification-free detection of DNA and RNA molecules in drawn blood samples. Similarly, He et  al. reported the use of chemiluminescence detection of alkaline phosphatase–streptavidin capable of binding superparamagnetic IONPs modified with biotin-labeled hepatitis B virus DNA [106]. In this particular case, the functionalization of MNPs with (3-aminopropyl) triethoxysilane (APTES) allowed further treatment with succinic anhydride to efficiently bind the biotinylated DNA.

4 Biomedical and Technological Applications of MNPs 4.1 Magnetic Separation of DNA Separation technology is one of the most complex and important areas of biotechnology research [107]. It is considered to be the starting point for downstream processes and product development and comprises molecular routine biotechnological activities such as DNA sequencing, amplification, cloning, and bio-detection [108]. Generally, successful nucleic acid purification requires four important steps: (1) effective disruption of cells or tissue; (2) denaturation of nucleoprotein complexes; (3) inactivation of nucleases (e.g., DNase for DNA extraction); and (4) secure storage away from any possibility of contamination. Traditional extraction and DNA purification techniques can be divided into two fundamental categories, namely, those in which purification is mediated with organic solvents and those in which purification occurs by solid-phase methodologies. Among those techniques in which purification is mediated with organic solvents, the guanidinium thiocyanate–phenol–chloroform extraction technique is considered to be a conventional method that comprises the formation of a biphasic emulsion to purify DNA. In short, two layers are formed by centrifugation, with one layer, a mixture of phenol–chloroform, used to internalize proteins, carbohydrates, and cell debris, and the second layer, the aqueous phase, containing the purified DNA molecules. The DNA is then precipitated using ethanol or isopropanol in 2:1 or 1:1 ratio and a high salts concentration [107]. Similar techniques, such as alkaline extraction Reprinted from the journal

29

13



Topics in Current Chemistry (2020) 378:13

[108] or cesium chloride gradient centrifugation [109], follow almost the same principles. In comparison to the above-mentioned convential methods, solid-phase isolation protocols allow a quick and efficient purification due to the prevention of incomplete phase separation in the liquid–liquid extraction and the use of toxic organic solvents. Solid-phase purification is most commonly performed using a spin column operated under centrifugal force. Silica matrices, glass particles, diatomaceous earth, and anion-exchange adsorbents are materials that have been employed as support in such systems [110]. The initial step in a solid-phase extraction process is to condition the column for sample adsorption. Column conditioning can be accomplished by using a buffer at a particular pH to convert the surface or functional groups on the solid column material into a particular chemical form. For the elution step, Tris–EDTA buffer or water is introduced to release the desired nucleic acid from the column so that it can be collected in a purified state [111]. As mentioned above, commonly used quick DNA isolation methods often contain chemicals that can lead to the degradation of DNA or be toxic to both humans and the environment [112, 113]. Moreover, they include time-consuming and laborintensive complex steps, such as centrifugation, precipitation, and filtration, all of which are able to compromise DNA integrity. A number of publications have reported in detail on DNA isolation techniques, emphasizing the special advantages of DNA magnetic separation [113, 114], which are a fast and simple handling of samples and the opportunity to deal with large volumes without the need for centrifugation steps. In addition, biomagnetic separation offers many benefits, including a high-quality product, simple treatment methodology, reduced need for chemicals, high-throughput system, and the potential for being used in automated processes [115]. The use of magnetic carriers such as IONPs functionalized with affinity ligands is preferred due to the high surface area and binding capacity of the IONPs and the ease manipulation [17]. Although many magnetic carriers are commercially available, the cost of using commercialized separation kits hinders the routine application of this facile technology for biochemical or clinical screening. In addition, new alternatives are developed almost every year in the continuing effort to improve separation efficiency. Such materials are MNPs modified with synthetic and natural polymers, porous glass, or material simply based on inorganic coatings, such as silica and organosilane precursors [17, 95, 116]. For example, Biao et al. reported the rapid purification of plasmid DNA from crude cell lysates using IONPs modified with silica [114]. These authors compared their method with a commercial kit as well as with a traditional phenol–chloroform technique, with their results demonstrating the advantages of their system. Tanaka et al. studied the adsorption and desorption behavior of DNA on aminosilane-modified MNPs for PCR analysis [117]. A comparison of some of the reported adsorbents in DNA isolation systems is given in Table 2. The successful application of silica-coated IONPs in DNA separation is associated with selective binding. Following the same principles as those for silica spin columns, such magnetic nanoplatforms require the use of a binding buffer to charge the surface of the nanoparticles. This surface charging creates a high affinity of the

13

30

Reprinted from the journal

Reprinted from the journal

31

UV/VIS

UV/VIS 10 min

30 min

24 extractions in 40 min

10 min

60 min

3 min

3 min

3 min

Desorption time

~ 98

> 96

83.4

86.16

~ 90

~ 85

~ 84

~ 73

Desorption percentage (%)

[95]

[27]

[27]

[118]

[117]

[52]

[52]

[52]

References

 Some features and comparisons of commercially kits based on magnetic particles used for DNA, RNA, and pDNA separation are described in detail by Berensmeier et al. [116]

b

 1-Hexyl-3-methylimidazolium bromide

a

UV, Ultraviolet; VIS, visible light spectrum

Silica and chitosan

[C6MIM]-Br

NUCLISENS®; EASYMAG®b

UV/VIS

Agarose gel analysis

M-MSN

Dimercaptosuccinic acid

a

UV/VIS

Silica

Silica

UV/VIS

UV/VIS

Tris(hydroxymethyl)aminomethane

Chitosan

Method

Coating agent

Table 2  Adsorption-related information of functionalized iron oxide nanoparticles during DNA separation

Topics in Current Chemistry (2020) 378:13  

13

Topics in Current Chemistry (2020) 378:13

negatively charged DNA backbone towards the positively charged interface. Sodium also plays a role as a cation bridge that attracts the negatively charged oxygen in the phosphate backbone of nucleic acids. As a chaotropic agent, sodium cations at high concentrations also promote the disruption of ordered water molecules along the DNA [119]. Since the nucleic acid is tightly bound, washings remove all contamination, and the purified DNA molecules can subsequently be eluted under low ionic strength using TE buffer or distilled water. In all of these cases no centrifugation steps are necessary due to the magnetic properties of the nanoparticle cores, and this feature is precisely the advantage of the methodology depicted in Fig. 6. The above-mentioned electrostatic interactions between MNPs and DNA must be taken into account and, consequently, the isolation protocol should be carefully designed. In general, electrostatic interactions between positively charged adsorbents and DNA molecules are considered to be the best alternative in terms of time and resources [52]. This protocol is not only ascribed for silica but also for a number of other functional groups, such as amino, hydroxyl, among others. Therefore, the use of binding and elution buffer is central to the separation procedure. As represented in Fig.  6, the DNA adsorption and elution stages can be mediated with acid–base buffer solutions to manipulate DNA recovery via pH switching [120]. 4.2 Non‑Viral DNA Delivery Systems: Magnetofection and Preparation of DNA Vaccines The extensive study of the DNA molecule in past decades has provided not only a better understanding of the fundamental basis of human life but also marked the beginning of the development of a novel group of therapies and diagnostic models. The application of gene therapy (or transfection) is currently receiving much attention due to the vast potential of such therapies. Basically, the main principle of gene therapy relies on the use of DNA as a pro-drug that can lead to the expression of therapeutic proteins within specific cells. Such DNA-based therapeutics include plasmids containing transgenes, oligonucleotides for antisense and antigen applications, ribozymes, aptamers, and small interfering RNAs (siRNAs) [121, 122]. Most

Fig. 6  Schematic procedure for nucleic acid purification using magnetic nanoparticles (NPs) as support

13

32

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13

of these are considered to be promising candidates for the treatment of a wide range of diseases, including cancer, human immunodeficiency virus, neurological malfunctions such as Parkinson’s disease and Alzheimer’s disease, and cardiovascular disorders [123]. The efficiency of transfection depends on the delivery of therapeutic agents to cells, with the consequence of effecting alterations in gene expression by replacing or silencing defective genetic material. In this process, a DNA carrier, commonly called as “transfectant,” is needed for the transportation of genes to the cellular compartment. The use of novel nucleic acid delivery systems has not only improved the pharmacokinetics of DNA-based therapeutics but also achieved an efficiently targeted introduction of these molecules into desired tissues and cells. DNA delivery techniques are currently classified into three general types, with varying efficacy [121, 124]: (1) stimuli-mediated techniques; (2) mechanical transfection, (3) and vector-assisted delivery systems. The first two types involve introducing naked DNA into cells via microinjection [125], photoporation, particle bombardment, sonoporation, or electroporation [126, 127], all of which are considered to be invasive methodologies. Such techniques are very precise, but they are time-consuming and restricted to local delivery in specific areas; moreover, they are rapidly degraded by serum nucleases. In contrast, the application of vector-assisted delivery systems is a suitable option for use in clinical trials. Viral vectors are currently the most effective gene delivery methodology (80–90%), but they are associated with the potential risk of inserting viral nucleic acid sequences into the host genome and potentially causing unwelcome effects, such as the inappropriate expression of genes. Consequently, safety, and immunogenicity concerns limit their usage in the current clinical scenario [128]. Non-viral vectors have important advantages over viral approaches due to their demonstrated biosafety in reducing pathogenicity, low cost, and ease of production. However, this approach is hindered by a lack of efficiency [127]. Non-viral gene therapy, within its broader context, includes such nucleic acid delivery applications as, for example, anti-sense or siRNAs, but the techniques used for nucleic acid delivery do not fall within the scope of this review, which we have limited to a specific focus on other procedures and materials used as non-viral vectors. As early as around 1990, plasmid DNA (pDNA) has been recognized for having an enormous potential for applications in gene therapy. Compared to viral and RNAbased vectors, plasmids are easier and cheaper to produce and store, and they have a much longer shelf life [127–129]. At the molecular level, plasmids employ the DNA transcription and translation apparatus in the cell to biosynthesize the therapeutic entity, namely, the protein. Thus, they are able to correct genetic errors that basically produce functionally incompetent copies of a given protein. In addition to focusing on a high molecular weight double-stranded DNA structure, several research groups have studied plasmid design [130, 131], with a special focus on the choice of enhancer, which is the pDNA region(s) that improve production of the targeted gene. Plasmids have also powered a large number of clinical trials as part of gene therapy in monogenic and polygenic diseases, such as cystic fibrosis and cancer, and in infectious diseases [131]. However, the selection of pDNA carrier and the route to the cell nucleus are possibly the two most challenging issues during transfection. Reprinted from the journal

33

13

Topics in Current Chemistry (2020) 378:13

Non-viral vector platforms or carriers should fulfill four important requirements: (1) capability to complex nucleic acids and protect them against nuclease enzymes at the extracellular compartment; (2) a positive net electric surface charge at physiological pH to overcome the negative potential of the cell membrane, since otherwise the cell membrane hinders the incorporation of negatively charged phosphate-containing DNA; (3) a mechanism to protect DNA from the acidic environment inside endosomes; and (4) chemical stability to maintain the integrity until the nucleus is reached [132]. Magnetofection techniques have been introduced in an attempt to fulfill these four important requirements and address the transient damage caused by the invasive methods mentioned above (i.e., microinjection, electroporation, among others). Magnetofection techniques are excellent alternative procedures that can significantly reduce the transfection time from several hours to < 60  min [28, 30, 132]. The association of superparamagnetic nanoparticles with gene vectors facilitates the transfection process into cells through the application of an external magnetic field that both targets and reduces the duration of the gene delivery, thereby enhancing the efficiency of the DNA vector (Fig. 7). The coating material is a key aspect of carrier design since these structures are responsible for DNA interaction as well as for DNA protection and chemical stabilization. Numerous materials have been used as coating agents for superparamagnetic IONPs, including cationic (bio)polymers, dendrimers, and cationic lipids (liposomal magnetofection). For example, Sohrabijam et  al. reported the use of chitosan-modified IONPs as a magnetofection carrier. Their results suggest a potentially enhanced magnetofection efficiency due to the cationic surface of the chitosan–IONPs [133]. Another example is the functionalization with polyethylenimine (PEI), which is considered to be one of the most interesting coating agents used in magnetofection due to its enormous stability [134]. The use of cationic lipids, such as N,N-di-n-hexadecyl-N,N-dihydroxyethylammonium chloride, has also been described to enhance DNA uptake in carcinogenic cells as

Fig. 7  Magnetofection process. Nuclear access is hampered first by the cell membrane, the cytoplasmic environment, and the nuclear membrane

13

34

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13

part of liposome formation [135]. Various examples of MNP applications in gene delivery that have been reported in the literature are shown in Table 3. Plasmids can be used not only as disease treatment, but also as a vaccination system for genetic immunization. A DNA vaccine is a third-generation vaccine that incorporates a vector with a eukaryotic cell promoter and a gene that encodes for an immunogenic protein. In contrast with protein-based traditional vaccines, DNA vaccination has the capacity to induce both cellular and humoral immune responses. The structure of plasmid DNA provides some advantages over other traditional protein-based or carbohydrate-based vaccines since plasmid DNA can encode many immunogenic proteins of the same virus and can also encode similar proteins belonging to different infective agents [145]. Other important advantages of DNA vaccines are their easy assembly, stability at room temperature, ease of manipulation, and low cost of plasmid production. However, the remaining challenges are to increase transfection efficiencies and facilitate intracellular uptake and to improve targeting to cells, as well as to perform these operations with small amounts of DNA. Various gene delivery and adjuvant systems at the nanoscale are used to overcome these problems. During the use of a nanotechnological adjuvant, the degradation of DNA is prevented, resulting in ultra-rapid delivery by targeting to the desired cells [146]. Adjuvants are generally grouped into two subtypes, namely, molecular adjuvants, which are immunostimulants, and carrier structures, which are systems that control release (e.g., mineral salts, liposomes, biodegradable polymers, and micro/nanoparticles). The use of cationic polymers or (polications) is associated with important advantages during transfection due to the possibility of electrostatic interactions with DNA molecules [93, 95, 133]. This property together with the superparamagnetic behavior of IONPs make these nanostructures excellent adjuvants for DNA vaccines. It has been reported that polymers bind with nucleic acid(s) to form complex structures known as polyplexes and that these polyplexes have increased transfection efficiency. Some of the examples reported in the literature include chitosan, PEI, poly(2-hydroxyethyl methacrylate) (pHEMA), polyamidoamine (PAMAM) dendrimers, polyethylene glycol (PEG), and poly-l-lysine [133, 145]. For example, Al-Deen et al. reported the use of IONPs/PEI/DNA polyplexes to enhance the delivery of a malaria DNA vaccine using magnetofection [147]. Their results indicate that in  vitro transfection efficiency into eukaryotic cells can be significantly enhanced under the application of an external magnetic field. Garu et  al. also described a novel DNA carrier based on lipoplexes of a model DNA vaccine using antibodylabeled MNPs [148]. This system revealed remarkable in vivo targeting properties of the described liposomal DNA vaccine carrier. Importantly, mice immunization induced a long-lasting anti-melanoma immune response. 4.3 Molecular Recognition Tools One of the advantages of using MNPs in molecular platform technologies is related to their easy manipulation due to a high separation efficiency with magnetic fields. The principal disadvantage is related to the aggregation process that could occur Reprinted from the journal

35

13

13

36

 PEG-block-poly(propylene glycol)-block-PEG

 Rat malignant glioma

 Human lung adenocarcinoma cell line

i

h

 Chinese hamster ovary cells

g

 Murine mammary adenocarcinoma

f

 Human umbilical vein endothelial cells

 Mouse embryonic carcinoma cells

e

d

 Human prostate carcinoma cells

c

b

 Poly(hexamethylene biguanide)

a

Plasmid

Plasmid

Poly(propyleneimine) dendrimers

Pluronic F-127b

Plasmid

Plasmid

Poly-l-Lysine

PEI (molecular weight: 25 K), chitosan

Plasmid

Polyacrylic acid and PEI siRNA

Plasmid

Hydroxyapatite

Branched PEI and ­PHMBGa

Plasmid

Deacylated polyethylenimine

H441i

Saos-2 osteoblasts

C6h

CHO-K1g and HeLa cells

Lung tissue

B16Ff

Rat marrow stromal cells

P19CL6e

37

12

45

80–90

60

47

60–70

80

80

23

HUVECd

DU145c

siRNA Plasmid

Polydopamine

Transfection efficiency (%)

Target cell

Vector

PEG, Branch PEI (molecular weight: 25 K)

Coating material

Table 3  Summary of some in situ magnetofection examples described in the literature using ion oxide nanoparticles as support

*

75

100

90

Not reported

Not reported

100

100

80

62.4

Cell viability (%)

[144]

[132]

[143]

[142]

[141]

[140]

[139]

[138]

[137]

[136]

References

Topics in Current Chemistry (2020) 378:13

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13



and consequently affect the sensitivity and specificity of the analysis. In this regard, molecular recognition involves specific noncovalent phenomena, such as hydrogen bonding, van der Waals, and hydrophobic forces, coordination to metals ions, π–π stacking, electrostatic or magnetic interactions, among others. These interactions play an important role in biological systems, as well as in the designing of biosensors for the early diagnosis of diseases and for therapeutic treatments by coupling with MNPs. For example, the double helix structure of DNA is found to be very stable due to some of these forces; specifically the Watson–Crick type of hydrogen bonds (guanine–cytosine and adenine–thymine). This type of interaction allows the development of specific nucleic acid probes, such as aptamers, for the molecular recognition of a wide range of targets comprising small molecules, proteins, and cells [26, 42, 90]. The development of new approaches for the selective detection of nucleic acids is currently one of the main challenges of the scientific community. Given the DNA sequence mismatches and structural folding of ssODN, strategies need to be developed that allow the detection of trace levels of specific sequences. The molecular recognition capacity of systems using ssODNs is a very sensitive and very specific compared with traditional recognition systems. In 1996, Tyagi and Kramer reported, for the first time, a new probe to detect specific nucleic acids in homogeneous solutions and also introduced the term ‘Molecular Beacon’ based on the fluorescence resonance energy transfer pair [149]. The single-stranded nucleic acid or hairpin molecules possess a stem-and-loop structure. The loop portion of the molecule is a probe sequence that binds a target nucleic acid sequence, while the stem is two complementary arm sequences that are annealed. The target and the arm sequences are not complementary. Fluorescent and non-fluorescent quenching moieties are attached to the end of both arms. When the target sequence is present, hybridization occurs and fluorescence is restored (Fig. 8). Consequently, the method is useful for detecting specific sequences of nucleic acids [149]. Molecular beacons can be applied in different fields, such as for the measurement of single nucleotide polymorphisms (genetic variations) [150–152], real-time

Fig. 8  Schematic representation of molecular beacon assembled to the core-shell IONPs@Au Reprinted from the journal

37

13



Topics in Current Chemistry (2020) 378:13

PCR quantification [153], multiple PCR assays [154], and clinical diagnosis [153, 155]. Another probe is the TaqMan assay, which is used to quantify mRNA levels of selected genes. Two fluorescent moieties with different wavelength emissions appear in the system. Some recent reports have included the use of MNPs as platforms for TaqMan probes. For example, Liu et al. described a portable quantitative and selective DNA detection biosensor. In only one step, target recognition occurs as a consequence of the releasing of the invertase–DNA conjugate which can be collected with the help of a magnet. The released invertase–DNA was used to catalyze the hydrolysis of sucrose into glucose with highly efficient sequence selectivity [156]. DNA-assembled core–satellite superstructures are other platforms used in drug delivery, imaging, and biosensing. Tian et al. reported an on-particle rolling circle amplification process [157] that allows rapid microRNA detection. Once the targeted microRNA is detected, the long ssDNA produced acts as the scaffold of the core–satellite superstructure and it can be hydrolyzed by duplex-specific nuclease. Due to hydrolyzation, MNPs are released and subsequently quantified in an optomagnetic sensor. The high capacity to discriminate single-nucleotide mismatches opens new possibilities to use these structures in clinical applications. The development of nanotechnology and its applications in the field of biotechnology lead to an improvement in bioanalysis. The bar-code analysis, which involves the use of IONPs and gold nanoparticles, is a diagnostic tool used for the detection of nucleic acids and proteins. Both nanoparticles are modified with recognition units that interact with the analyte to form a sandwich-type structure. The magnetic properties of the designed system can then be used to separate the sandwich structures [158, 159]. Gold nanoparticles can also be functionalized with a shell of barcode consisting of hybridized oligonucleotides. After the separation of the sandwich structure, the strands of bar codes can be identified in the microarrays by PCR or another analytical tool. In addition, there has been a recent report of a method for the detection of gastric cancer in which DNA probes, magnetic nanoprobes, and silicon–gold nanoparticles are hybridized to form a sandwich structure; the conjugate is then magnetically separated and the Au–nanoparticle released [160]. The authors suggest that complementary and mismatched DNA can be clearly distinguished by using inductively coupled plasma mass spectrometry in the detection of DNA with high sensitivity and specificity.

5 Conclusions The systematic use of DNA-based magnetic nanoplatforms shows great potential in biotechnological fields due to the important advantages of these platforms in terms of time and resources. In this regard, relatively cheap and high-throughput DNA extraction procedures have overcome the main drawbacks associated with traditional methodologies. Compared with the well-known chemical recognition mechanisms, such as host–guest chemistry, the interaction of nucleic acids is universal and easily modified. As a consequence, several biosensors and diagnostic methods have been

13

38

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13



developed based on the magnetic behavior of MNPs and the special recognition properties of DNA. The appropriate selection of the delivery vector is a key factor in drug and gene delivery systems. DNA molecules and their interaction with the surface of MNPs should be carefully designed taking into acount the the final purpose and all the aspects related to the host cell environment. Future perspectives in vaccination routines will probably include DNA vaccine involving nanoparticles as a novel method to generate antigen-specific antibodies and cell-mediated immunity. However, many aspects still need to be optimized, such as the nanoparticle–DNA vaccine efficiency and cell viability, together with their complete clinical assays in human subjects to validate immunogenicity. Compliance with ethical standards  Conflict of interest  The authors declare no conflict of interest.

References 1. Jeevanandam J, Barhoum A, Chan YS et  al (2018) Review on nanoparticles and nanostructured materials: history, sources, toxicity and regulations. Beilstein J Nanotechnol 9:1050–1074. https​:// doi.org/10.3762/bjnan​o.9.98 2. Ahsan MA, Jabbari V, Islam MT et  al (2019) Sustainable synthesis and remarkable adsorption capacity of MOF/graphene oxide and MOF/CNT based hybrid nanocomposites for the removal of Bisphenol A from water. Sci Total Environ 673:306–317. https​://doi.org/10.1016/j.scito​ tenv.2019.03.219 3. Ahsan MA, Jabbari V, Imam MA et al (2020) Nanoscale nickel metal organic framework decorated over graphene oxide and carbon nanotubes for water remediation. Sci Total Environ 698:134214. https​://doi.org/10.1016/j.scito​tenv.2019.13421​4 4. Ahsan MA, Deemer E, Fernandez-Delgado O et al (2019) Fe nanoparticles encapsulated in MOFderived carbon for the reduction of 4-nitrophenol and methyl orange in water. Catal Commun 130:105753. https​://doi.org/10.1016/j.catco​m.2019.10575​3 5. Ahsan MA, Fernandez-Delgado O, Deemer E et al (2019) Carbonization of Co-BDC MOF results in magnetic C@Co nanoparticles that catalyze the reduction of methyl orange and 4-nitrophenol in water. J Mol Liq 290:111059. https​://doi.org/10.1016/j.molli​q.2019.11105​9 6. Ahsan MA, Jabbari V, El-Gendy AA et  al (2019) Ultrafast catalytic reduction of environmental pollutants in water via MOF-derived magnetic Ni and Cu nanoparticles encapsulated in porous carbon. Appl Surf Sci 497:143608. https​://doi.org/10.1016/j.apsus​c.2019.14360​8 7. Franco A, Cebrián-García S, Rodríguez-Padrón D et al (2018) Encapsulated laccases as effective electrocatalysts for oxygen reduction reactions. ACS Sustain Chem Eng 6:11058–11062. https​:// doi.org/10.1021/acssu​schem​eng.8b025​29 8. Cova CM, Zuliani A, Puente Santiago AR et al (2018) Microwave-assisted preparation of Ag/Ag2S carbon hybrid structures from pig bristles as efficient HER catalysts. J Mater Chem A 6:21516– 21523. https​://doi.org/10.1039/C8TA0​6417B​ 9. Ostovar S, Franco A, Puente-Santiago AR et  al (2018) Efficient mechanochemical bifunctional nanocatalysts for the conversion of isoeugenol to vanillin. Front Chem 6:1–7. https​://doi. org/10.3389/fchem​.2018.00077​ 10. Rodríguez-Padrón D, Puente-Santiago AR, Balu AM et al (2019) Environmental catalysis: present and future. ChemCatChem 11:18–38. https​://doi.org/10.1002/cctc.20180​1248 11. Feng S, Li D, Low Z et  al (2017) ALD-seeded hydrothermally-grown Ag/ZnO nanorod PTFE membrane as efficient indoor air filter. J Memb Sci 531:86–93. https​://doi.org/10.1016/j.memsc​ i.2017.02.042 Reprinted from the journal

39

13



Topics in Current Chemistry (2020) 378:13

12. Shalan AE, El-Shazly AN, Rashad MM, Allam NK (2019) Tin-zinc-oxide nanocomposites (SZO) as promising electron transport layers for efficient and stable perovskite solar cells. Nanoscale Adv 1:2654–2662. https​://doi.org/10.1039/c9na0​0182d​ 13. Sanad MF, Shalan AE, Bazid SM et al (2019) A graphene gold nanocomposite-based 5-FU drug and the enhancement of the MCF-7 cell line treatment. RSC Adv 9:31021–31029. https​://doi. org/10.1039/C9RA0​5669F​ 14. León Félix L, Sanz B, Sebastián V et al (2019) Gold-decorated magnetic nanoparticles design for hyperthermia applications and as a potential platform for their surface-functionalization. Sci Rep 9:4185. https​://doi.org/10.1038/s4159​8-019-40769​-2 15. Saif S, Tahir A, Asim T et  al (2019) Polymeric nanocomposites of iron-oxide nanoparticles (IONPs) synthesized using Terminalia chebula leaf extract for enhanced adsorption of arsenic(V) from water. Colloids Interfaces 3:17. https​://doi.org/10.3390/collo​ids30​10017​ 16. Kim H-M, Kim D, Jeong C et al (2018) Assembly of plasmonic and magnetic nanoparticles with fluorescent silica shell layer for tri-functional SERS-magnetic-fluorescence probes and its bioapplications. Sci Rep 8:13938. https​://doi.org/10.1038/s4159​8-018-32044​-7 17. Pershina AG, Sazonov AE, Filimonov VD (2014) Magnetic nanoparticles—DNA interactions: design and applications of nanobiohybrid systems. Russ Chem Rev 83:299–322. https​://doi. org/10.1070/RC201​4v083​n04AB​EH004​412 18. Feng Q, Liu Y, Huang J et al (2018) Uptake, distribution, clearance, and toxicity of iron oxide nanoparticles with different sizes and coatings. Sci Rep 8:2082. https​://doi.org/10.1038/s4159​8-01819628​-z 19. Huerta-Nuñez LFE, Gutierrez-Iglesias G, Martinez-Cuazitl A et  al (2019) A biosensor capable of identifying low quantities of breast cancer cells by electrical impedance spectroscopy. Sci Rep 9:6419. https​://doi.org/10.1038/s4159​8-019-42776​-9 20. Wei Y, Liao R, Mahmood AA et al (2017) pH-responsive pHLIP (pH low insertion peptide) nanoclusters of superparamagnetic iron oxide nanoparticles as a tumor-selective MRI contrast agent. Acta Biomater 55:194–203. https​://doi.org/10.1016/j.actbi​o.2017.03.046 21. Sahoo SL, Liu C-H (2015) Adsorption behaviors of DNA by modified magnetic nanoparticles: effect of spacer and salt. Colloids Surf A Physicochem Eng Asp 482:184–194. https​://doi. org/10.1016/j.colsu​rfa.2015.05.010 22. Haddad Y, Xhaxhiu K, Kopel P et al (2016) The isolation of DNA by polycharged magnetic particles: an analysis of the interaction by zeta potential and particle size. Int J Mol Sci 17:550. https​:// doi.org/10.3390/ijms1​70405​50 23. Robinson I, Tung LD, Maenosono S et  al (2010) Synthesis of core–shell gold coated magnetic nanoparticles and their interaction with thiolated DNA. Nanoscale 2:2624. https​://doi.org/10.1039/ c0nr0​0621a​ 24. Esmaeili E, Ghiass MA, Vossoughi M, Soleimani M (2017) Hybrid magnetic-DNA directed immobilisation approach for efficient protein capture and detection on microfluidic platforms. Sci Rep 7:194. https​://doi.org/10.1038/s4159​8-017-00268​-8 25. Sun W, Fletcher D, van Heeckeren RC, Davis PB (2012) Non-covalent ligand conjugation to biotinylated DNA nanoparticles using TAT peptide genetically fused to monovalent streptavidin. J Drug Target 20:678–690. https​://doi.org/10.3109/10611​86X.2012.71212​8 26. Cheon HJ, Lee SM, Kim S-R et  al (2018) Colorimetric detection of MPT64 antibody based on an aptamer adsorbed magnetic nanoparticles for diagnosis of tuberculosis. J Nanosci Nanotechnol 19:622–626. https​://doi.org/10.1166/jnn.2019.15905​ 27. Ghaemi M, Absalan G (2014) Study on the adsorption of DNA on F ­ e3O4 nanoparticles and on ionic liquid-modified ­Fe3O4 nanoparticles. Microchim Acta 181:45–53. https​://doi.org/10.1007/ s0060​4-013-1040-5 28. Smolders S, Kessels S, Smolders SM-T et  al (2018) Magnetofection is superior to other chemical transfection methods in a microglial cell line. J Neurosci Methods 293:169–173. https​://doi. org/10.1016/j.jneum​eth.2017.09.017 29. Megías R, Arco M, Ciriza J et al (2017) Design and characterization of a magnetite/PEI multifunctional nanohybrid as non-viral vector and cell isolation system. Int J Pharm 518:270–280. https​:// doi.org/10.1016/j.ijpha​rm.2016.12.042 30. Singh J, Mohanty I, Rattan S (2018) In vivo magnetofection: a novel approach for targeted topical delivery of nucleic acids for rectoanal motility disorders. Am J Physiol Liver Physiol 314:G109– G118. https​://doi.org/10.1152/ajpgi​.00233​.2017

13

40

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13 31. Wang W, Wang Y, Tu L et al (2015) Magnetoresistive performance and comparison of supermagnetic nanoparticles on giant magnetoresistive sensor-based detection system. Sci Rep 4:5716. https​://doi.org/10.1038/srep0​5716 32. Elgqvist J (2017) Nanoparticles as theranostic vehicles in experimental and clinical applications— focus on prostate and breast cancer. Int J Mol Sci 18:1102. https​://doi.org/10.3390/ijms1​80511​02 33. Sungsuwan S, Yin Z, Huang X (2015) Lipopeptide-coated iron oxide nanoparticles as potential glycoconjugate-based synthetic anticancer vaccines. ACS Appl Mater Interfaces 7:17535–17544. https​://doi.org/10.1021/acsam​i.5b054​97 34. Ernst C, Bartel A, Elferink JW et al (2019) Improved DNA extraction and purification with magnetic nanoparticles for the detection of methicillin-resistant Staphylococcus aureus. Vet Microbiol 230:45–48. https​://doi.org/10.1016/j.vetmi​c.2019.01.009 35. Sharma A, Goyal AK, Rath G (2018) Recent advances in metal nanoparticles in cancer therapy. J Drug Target 26:617–632. https​://doi.org/10.1080/10611​86X.2017.14005​53 36. Li S, Tang F, Wang H et al (2018) Au–Ag and Pt–Ag bimetallic nanoparticles@halloysite nanotubes: morphological modulation, improvement of thermal stability and catalytic performance. RSC Adv 8:10237–10245. https​://doi.org/10.1039/C8RA0​0423D​ 37. Mazrouaa AM, Mohamed MG, Fekry M (2019) Physical and magnetic properties of iron oxide nanoparticles with a different molar ratio of ferrous and ferric. Egypt J Pet 28:165–171. https​://doi. org/10.1016/j.ejpe.2019.02.002 38. Smith M, McKeague M, DeRosa MC (2019) Synthesis, transfer, and characterization of core– shell gold-coated magnetic nanoparticles. MethodsX 6:333–354. https​://doi.org/10.1016/j. mex.2019.02.006 39. Jishkariani D, Wu Y, Wang D et  al (2017) Preparation and self-assembly of dendronized Janus ­Fe3O4–Pt and ­Fe3O4–Au heterodimers. ACS Nano 11:7958–7966. https​://doi.org/10.1021/acsna​ no.7b024​85 40. Nikitin A, Khramtsov M, Garanina A et  al (2019) Synthesis of iron oxide nanorods for enhanced magnetic hyperthermia. J Magn Magn Mater 469:443–449. https​://doi.org/10.1016/j. jmmm.2018.09.014 41. Lv YB, Chandrasekharan P, Li Y et al (2018) Magnetic resonance imaging quantification and biodistribution of magnetic nanoparticles using T1-enhanced contrast. J Mater Chem B 6:1470–1478. https​://doi.org/10.1039/C7TB0​3129G​ 42. Hong L, Zhou F, Shi D et al (2017) Portable aptamer biosensor of platelet-derived growth factorBB using a personal glucose meter with triply amplified. Biosens Bioelectron 95:152–159. https​:// doi.org/10.1016/j.bios.2017.04.023 43. Farahbakhsh F, Ahmadi M, Hekmatara SH et al (2019) Improvement of photocatalyst properties of magnetic NPs by new anionic surfactant. Mater Chem Phys 224:279–285. https​://doi.org/10.1016/j. match​emphy​s.2018.11.074 44. Ivashchenko O, Peplińska B, Gapiński J et al (2018) Silver and ultrasmall iron oxides nanoparticles in hydrocolloids: effect of magnetic field and temperature on self-organization. Sci Rep 8:4041. https​://doi.org/10.1038/s4159​8-018-22426​-2 45. Demangeat E, Pédrot M, Dia A et al (2018) Colloidal and chemical stabilities of iron oxide nanoparticles in aqueous solutions: the interplay of structural, chemical and environmental drivers. Environ Sci Nano 5:992–1001. https​://doi.org/10.1039/C7EN0​1159H​ 46. Gupta R, Sharma D (2019) Biofunctionalization of magnetite nanoparticles with stevioside: effect on the size and thermal behaviour for use in hyperthermia applications. Int J Hyperth 36:302–312. https​://doi.org/10.1080/02656​736.2019.15657​87 47. Kurapov YA, Vazhnichaya EM, Litvin SE et  al (2019) Physical synthesis of iron oxide nanoparticles and their biological activity in  vivo. SN Appl Sci 1:102. https​://doi.org/10.1007/s4245​ 2-018-0110-z 48. Yazdani F, Seddigh M (2016) Magnetite nanoparticles synthesized by co-precipitation method: the effects of various iron anions on specifications. Mater Chem Phys 184:318–323. https​://doi. org/10.1016/j.match​emphy​s.2016.09.058 49. Maity D, Choo S-G, Yi J et  al (2009) Synthesis of magnetite nanoparticles via a solvent-free thermal decomposition route. J Magn Magn Mater 321:1256–1259. https​://doi.org/10.1016/j. jmmm.2008.11.013 50. Ansari S, Ficiarà E, Ruffinatti F et al (2019) Magnetic iron oxide nanoparticles: synthesis, characterization and functionalization for biomedical applications in the central nervous system. Materials (Basel) 12:465. https​://doi.org/10.3390/ma120​30465​ Reprinted from the journal

41

13

Topics in Current Chemistry (2020) 378:13 51. Obayemi JD, Dozie-Nwachukwu S, Danyuo Y et  al (2015) Biosynthesis and the conjugation of magnetite nanoparticles with luteinizing hormone releasing hormone (LHRH). Mater Sci Eng C 46:482–496. https​://doi.org/10.1016/j.msec.2014.10.081 52. Sosa-Acosta J, Silva JA, Fernández-Izquierdo L et al (2018) Iron oxide nanoparticles (IONPs) with potential applications in plasmid DNA isolation. Colloids Surf A Physicochem Eng Asp 545:167– 178. https​://doi.org/10.1016/j.colsu​rfa.2018.02.062 53. LaGrow AP, Besenhard MO, Hodzic A et al (2019) Unravelling the growth mechanism of the coprecipitation of iron oxide nanoparticles with the aid of synchrotron X-ray diffraction in solution. Nanoscale 11:6620–6628. https​://doi.org/10.1039/C9NR0​0531E​ 54. Laurent S, Forge D, Port M et  al (2008) Magnetic iron oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem Rev 108:2064–2110. https​://doi.org/10.1021/cr068​445e 55. Lassenberger A, Grünewald TA, van Oostrum PDJ et al (2017) Monodisperse iron oxide nanoparticles by thermal decomposition: elucidating particle formation by second-resolved in situ smallangle X-ray scattering. Chem Mater 29:4511–4522. https​://doi.org/10.1021/acs.chemm​ater.7b012​ 07 56. Cotin G, Kiefer C, Perton F et  al (2018) Unravelling the thermal decomposition parameters for the synthesis of anisotropic iron oxide nanoparticles. Nanomaterials 8:881. https​://doi.org/10.3390/ nano8​11088​1 57. Nam J-H, Joo Y-H, Lee J-H et  al (2009) Preparation of NiZn-ferrite nanofibers by electrospinning for DNA separation. J Magn Magn Mater 321:1389–1392. https​://doi.org/10.1016/j. jmmm.2009.02.044 58. Lee J-H, Huh Y-M, Jun Y et  al (2007) Artificially engineered magnetic nanoparticles for ultrasensitive molecular imaging. Nat Med 13:95–99. https​://doi.org/10.1038/nm146​7 59. Prodělalová J, Rittich B, Španová A et  al (2004) Isolation of genomic DNA using magnetic cobalt ferrite and silica particles. J Chromatogr A 1056:43–48. https​://doi.org/10.1016/j.chrom​ a.2004.08.090 60. Zheng J, Hu L, Zhang M et al (2015) An electrochemical sensing strategy for the detection of the hepatitis B virus sequence with homogenous hybridization based on host–guest recognition. RSC Adv 5:92025–92032. https​://doi.org/10.1039/C5RA1​6204A​ 61. Salehiabar M, Nosrati H, Davaran S et al (2018) Facile synthesis and characterization of l-aspartic acid coated iron oxide magnetic nanoparticles (IONPs) for biomedical applications. Drug Res (Stuttg) 68:280–285. https​://doi.org/10.1055/s-0043-12019​7 62. Goroncy C, Saloga PEJ, Gruner M et al (2018) Influence of organic ligands on the surface oxidation state and magnetic properties of iron oxide particles. Z Phys Chem 232:819–844. https​://doi. org/10.1515/zpch-2017-1084 63. Nosrati H, Salehiabar M, Davaran S et  al (2018) Methotrexate-conjugated l-lysine coated iron oxide magnetic nanoparticles for inhibition of MCF-7 breast cancer cells. Drug Dev Ind Pharm 44:886–894. https​://doi.org/10.1080/03639​045.2017.14174​22 64. Piotrowski P, Krogul-Sobczak A, Kaim A (2019) Magnetic iron oxide nanoparticles functionalized with C60 phosphonic acid derivative for catalytic reduction of 4-nitrophenol. J Environ Chem Eng 7:103147. https​://doi.org/10.1016/j.jece.2019.10314​7 65. Xu Y, Qin Y, Palchoudhury S, Bao Y (2011) Water-soluble iron oxide nanoparticles with high stability and selective surface functionality. Langmuir 27:8990–8997. https​://doi.org/10.1021/la201​ 652h 66. Sun M, Dai B, Liu K et  al (2018) Enhancement in thermal conductivity of polymer composites using aligned diamonds coated with superparamagnetic magnetite. Compos Sci Technol 164:129– 135. https​://doi.org/10.1016/j.comps​citec​h.2018.05.039 67. Fan Q, Guan Y, Zhang Z et  al (2019) A new method of synthesis well-dispersion and dense ­Fe3O4@SiO2 magnetic nanoparticles for DNA extraction. Chem Phys Lett 715:7–13. https​://doi. org/10.1016/j.cplet​t.2018.11.001 68. Hufschmid R, Teeman E, Mehdi BL et  al (2019) Observing the colloidal stability of iron oxide nanoparticles in situ. Nanoscale 11:13098–13107. https​://doi.org/10.1039/C9NR0​3709H​ 69. Schroffenegger M, Reimhult E (2018) Thermoresponsive core–shell nanoparticles: does core size matter? Materials (Basel) 11:1654. https​://doi.org/10.3390/ma110​91654​ 70. Iriarte-Mesa C, Díaz-Castañón S, Abradelo DG (2019) Facile immobilization of Trametes versicolor laccase on highly monodisperse superparamagnetic iron oxide nanoparticles. Colloids Surf B Biointerfaces 181:470–479. https​://doi.org/10.1016/j.colsu​rfb.2019.05.012

13

42

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13



71. Du L, Wang W, Zhang C et  al (2018) A versatile coordinating ligand for coating semiconductor, metal, and metal oxide nanocrystals. Chem Mater 30:7269–7279. https​://doi.org/10.1021/acs. chemm​ater.8b035​27 72. Veisi H, Razeghi S, Mohammadi P, Hemmati S (2019) Silver nanoparticles decorated on thiolmodified magnetite nanoparticles ­(Fe3O4/SiO2-Pr-S-Ag) as a recyclable nanocatalyst for degradation of organic dyes. Mater Sci Eng C 97:624–631. https​://doi.org/10.1016/j.msec.2018.12.076 73. Ebrahiminezhad A, Ghasemi Y, Rasoul-Amini S et al (2012) Impact of amino-acid coating on the synthesis and characteristics of iron-oxide nanoparticles (IONs). Bull Korean Chem Soc 33:3957– 3962. https​://doi.org/10.5012/bkcs.2012.33.12.3957 74. Nosrati H, Salehiabar M, Davaran S et al (2017) New advances strategies for surface functionalization of iron oxide magnetic nano particles (IONPs). Res Chem Intermed 43:7423–7442. https​://doi. org/10.1007/s1116​4-017-3084-3 75. Bai Y, Roncancio D, Suo Y et al (2019) A method based on amino-modified magnetic nanoparticles to extract DNA for PCR-based analysis. Colloids Surf B Biointerfaces 179:87–93. https​://doi. org/10.1016/j.colsu​rfb.2019.03.005 76. Oza G, Krishnajyothi K, Merupo VI et al (2019) Gold-iron oxide yolk-shell nanoparticles (YSNPs) as magnetic probe for fluorescence-based detection of 3 base mismatch DNA. Colloids Surf B Biointerfaces 176:431–438. https​://doi.org/10.1016/j.colsu​rfb.2019.01.016 77. Chen WD, Kohll AX, Nguyen BH et al (2019) Combining data longevity with high storage capacity—layer-by-layer DNA encapsulated in magnetic nanoparticles. Adv Funct Mater 29:1901672. https​://doi.org/10.1002/adfm.20190​1672 78. Wang L, Yao M, Fang X, Yao X (2019) Novel competitive chemiluminescence DNA assay based on ­Fe3O4@SiO2@Au-functionalized magnetic nanoparticles for sensitive detection of p53 tumor suppressor gene. Appl Biochem Biotechnol 187:152–162. https​://doi.org/10.1007/s1201​ 0-018-2808-1 79. Dalmina M, Pittella F, Sierra JA et  al (2019) Magnetically responsive hybrid nanoparticles for in  vitro siRNA delivery to breast cancer cells. Mater Sci Eng C 99:1182–1190. https​://doi. org/10.1016/j.msec.2019.02.026 80. Bakshi S, Zakharchenko A, Minko S et al (2019) Towards nanomaterials for cancer theranostics: a system of DNA-modified magnetic nanoparticles for detection and suppression of RNA marker in cancer cells. Magnetochemistry 5:24. https​://doi.org/10.3390/magne​toche​mistr​y5020​024 81. Khadsai S, Seeja N, Deepuppha N et al (2018) Poly(acrylic acid)-grafted magnetite nanoparticle conjugated with pyrrolidinyl peptide nucleic acid for specific adsorption with real DNA. Colloids Surf B Biointerfaces 165:243–251. https​://doi.org/10.1016/j.colsu​rfb.2018.02.039 82. Song J, Lei T, Yang Y et al (2018) Attachment of enzymes to hydrophilic magnetic nanoparticles through DNA-directed immobilization with enhanced stability and catalytic activity. New J Chem 42:8458–8468. https​://doi.org/10.1039/C8NJ0​0426A​ 83. Karami F, Noori-Daloii MR, Omidfar K et  al (2018) Modified methylated DNA immunoprecipitation protocol for noninvasive prenatal diagnosis of Down syndrome. J Obstet Gynaecol Res 44:608–613. https​://doi.org/10.1111/jog.13577​ 84. Ceylan Ş, Odabaşı M (2013) Novel adsorbent for DNA adsorption: F ­ e3+-attached sporopollenin particles embedded composite cryogels. Artif Cells Nanomed Biotechnol 41:376–383. https​://doi. org/10.3109/21691​401.2012.75912​5 85. Liu B, Liu J (2014) DNA adsorption by magnetic iron oxide nanoparticles and its application for arsenate detection. Chem Commun 50:8568. https​://doi.org/10.1039/C4CC0​3264K​ 86. Guo Y, Wang Y, Li S et al (2017) DNA-spheres decorated with magnetic nanocomposites based on terminal transfer reactions for versatile target detection and cellular targeted drug delivery. Chem Commun 53:4826–4829. https​://doi.org/10.1039/C7CC0​0310B​ 87. Wang H, Yang R, Yang L, Tan W (2009) Nucleic acid conjugated nanomaterials for enhanced molecular recognition. ACS Nano 3:2451–2460. https​://doi.org/10.1021/nn900​6303 88. Panda D, Saha P, Das T, Dash J (2017) Target guided synthesis using DNA nano-templates for selectively assembling a G-quadruplex binding c-MYC inhibitor. Nat Commun 8:16103. https​:// doi.org/10.1038/ncomm​s1610​3 89. Stanciu L, Won Y-H, Ganesana M, Andreescu S (2009) Magnetic particle-based hybrid platforms for bioanalytical sensors. Sensors 9:2976–2999. https​://doi.org/10.3390/s9040​2976 90. Tintoré M, Mazzini S, Polito L et al (2015) Gold-coated superparamagnetic nanoparticles for single methyl discrimination in DNA aptamers. Int J Mol Sci 16:27625–27639. https​://doi.org/10.3390/ ijms1​61126​046 Reprinted from the journal

43

13



Topics in Current Chemistry (2020) 378:13

91. Slavin S, De Cuendias A, Ladmiral V, Haddleton DM (2011) Biotin functionalized poly(sulfonic acid)s for bioconjugation: in  situ binding monitoring by QCM-D. J Polym Sci A Polym Chem 49:1163–1173. https​://doi.org/10.1002/pola.24532​ 92. Trigueros Domènech, Toulis Marfany (2019) In vitro gene delivery in retinal pigment epithelium cells by plasmid DNA-wrapped gold nanoparticles. Genes (Basel) 10:289. https​://doi.org/10.3390/ genes​10040​289 93. Pandit KR, Nanayakkara IA, Cao W et al (2015) Capture and direct amplification of DNA on chitosan microparticles in a single PCR-optimal solution. Anal Chem 87:11022–11029. https​://doi. org/10.1021/acs.analc​hem.5b030​06 94. Liu Y, Li Y, Li X-M, He T (2013) Kinetics of (3-aminopropyl)triethoxylsilane (APTES) silanization of superparamagnetic iron oxide nanoparticles. Langmuir 29:15275–15282. https​://doi. org/10.1021/la403​269u 95. Tiwari AP, Satvekar RK, Rohiwal SS et  al (2015) Magneto-separation of genomic deoxyribose nucleic acid using pH responsive ­Fe3O4 @silica@chitosan nanoparticles in biological samples. RSC Adv 5:8463–8470. https​://doi.org/10.1039/C4RA1​5806G​ 96. Bui TQ, Ngo HTM, Tran HT (2018) Surface-protective assistance of ultrasound in synthesis of superparamagnetic magnetite nanoparticles and in preparation of mono-core magnetite-silica nanocomposites. J Sci Adv Mater Devices 3:323–330. https​://doi.org/10.1016/j.jsamd​.2018.07.002 97. Li Z, Chen H, Bao H, Gao M (2004) One-pot reaction to synthesize water-soluble magnetite nanocrystals. Chem Mater 16:1391–1393. https​://doi.org/10.1021/cm035​346y 98. Byrne SJ, Corr SA, Gun’ko YK et al (2004) Magnetic nanoparticle assemblies on denatured DNA show unusual magnetic relaxivity and potential applications for MRI. Chem Commun 10:2560. https​://doi.org/10.1039/b4096​03g 99. Mohamed HDA, Watson SMD, Horrocks BR, Houlton A (2012) Magnetic and conductive magnetite nanowires by DNA-templating. Nanoscale 4:5936. https​://doi.org/10.1039/c2nr3​1559a​ 100. Sreenivasulu G, Lochbiler TA, Panda M et al (2016) Self-assembly of multiferroic core–shell particulate nanocomposites through DNA–DNA hybridization and magnetic field directed assembly of superstructures. AIP Adv 6:045202. https​://doi.org/10.1063/1.49457​61 101. Zhu N, Zhang A, He P, Fang Y (2004) DNA hybridization at magnetic nanoparticles with electrochemical stripping detection. Electroanalysis 16:1925–1930. https​://doi.org/10.1002/elan.20030​ 3028 102. Wang F, Shen H, Feng J, Yang H (2006) PNA-modified magnetic nanoparticles and their hybridization with single-stranded DNA target: surface enhanced Raman scatterings study. Microchim Acta 153:15–20. https​://doi.org/10.1007/s0060​4-005-0460-2 103. Diamandis EP, Christopoulos TK (1991) The biotin-(strept)avidin system: principles and applications in biotechnology. Clin Chem 37:625–636 104. de Freitas CF, Montanha MC, Pellosi DS et  al (2019) Biotin-targeted mixed liposomes: a smart strategy for selective release of a photosensitizer agent in cancer cells. Mater Sci Eng C 104:109923. https​://doi.org/10.1016/j.msec.2019.10992​3 105. Cannon B, Campos AR, Lewitz Z et  al (2012) Zeptomole detection of DNA nanoparticles by single-molecule fluorescence with magnetic field-directed localization. Anal Biochem 431:40–47. https​://doi.org/10.1016/j.ab.2012.08.017 106. He N, Wang F, Ma C et al (2013) Chemiluminescence analysis for HBV-DNA hybridization detection with magnetic nanoparticles based DNA extraction from positive whole blood samples. J Biomed Nanotechnol 9:267–273. https​://doi.org/10.1166/jbn.2013.1478 107. Oberacker P, Stepper P, Bond DM et al (2019) Bio-On-Magnetic-Beads (BOMB): open platform for high-throughput nucleic acid extraction and manipulation. PLoS Biol 17:e3000107. https​://doi. org/10.1371/journ​al.pbio.30001​07 108. Chomczynski P, Sacchi N (2006) The single-step method of RNA isolation by acid guanidinium thiocyanate–phenol–chloroform extraction: twenty-something years on. Nat Protoc 1:581–585. https​://doi.org/10.1038/nprot​.2006.83 109. Delaney S, Murphy R, Walsh F (2018) A comparison of methods for the extraction of plasmids capable of conferring antibiotic resistance in a human pathogen from complex broiler cecal samples. Front Microbiol 9:1731. https​://doi.org/10.3389/fmicb​.2018.01731​ 110. Sermwittayawong D, Jakkawanpitak C, Waji N, Hutadilok-Towatana N (2013) Economical method for midiprep plasmid DNA purification using diatomaceous earth. ScienceAsia 39:631. https​://doi. org/10.2306/scien​ceasi​a1513​-1874.2013.39.631

13

44

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13 111. Bai Y, Cui Y, Suo Y et  al (2019) A rapid method for detection of salmonella in milk based on extraction of mRNA using magnetic capture probes and RT-qPCR. Front Microbiol. https​://doi. org/10.3389/fmicb​.2019.00770​ 112. Griffiths L, Chacon-Cortes D (2014) Methods for extracting genomic DNA from whole blood samples: current perspectives. J Biorepository Sci Appl Med 2:1. https​://doi.org/10.2147/BSAM. S4657​3 113. Di Pietro F, Ortenzi F, Tilio M et  al (2011) Genomic DNA extraction from whole blood stored from 15- to 30-years at −20°C by rapid phenol–chloroform protocol: a useful tool for genetic epidemiology studies. Mol Cell Probes 25:44–48. https​://doi.org/10.1016/j.mcp.2010.10.003 114. Neng-Biao W, Lian X, Li-na C et al (2009) Isolation and purification of plasmid D NA by silicacoated magnetic nanoparticles. Chin J Biochem Mol Biol 25:958–962 115. Köse K (2016) Nucleotide incorporated magnetic microparticles for isolation of DNA. Process Biochem 51:1644–1649. https​://doi.org/10.1016/j.procb​io.2016.07.021 116. Berensmeier S (2006) Magnetic particles for the separation and purification of nucleic acids. Appl Microbiol Biotechnol 73:495–504. https​://doi.org/10.1007/s0025​3-006-0675-0 117. Tanaka T, Sakai R, Kobayashi R et al (2009) Contributions of phosphate to DNA adsorption/desorption behaviors on aminosilane-modified magnetic nanoparticles. Langmuir 25:2956–2961. https​ ://doi.org/10.1021/la803​2397 118. Min JH, Woo M-K, Yoon HY et  al (2014) Isolation of DNA using magnetic nanoparticles coated with dimercaptosuccinic acid. Anal Biochem 447:114–118. https​://doi.org/10.1016/j. ab.2013.11.018 119. Smerkova K, Dostalova S, Vaculovicova M et al (2013) Investigation of interaction between magnetic silica particles and lambda phage DNA fragment. J Pharm Biomed Anal 86:65–72. https​:// doi.org/10.1016/j.jpba.2013.07.039 120. Saiyed ZM, Bochiwal C, Gorasia H et  al (2006) Application of magnetic particles ­(Fe3O4) for isolation of genomic DNA from mammalian cells. Anal Biochem 356:306–308. https​://doi. org/10.1016/j.ab.2006.06.027 121. Gessner I, Yu X, Jüngst C et al (2019) Selective capture and purification of microRNAs and intracellular proteins through antisense-vectorized magnetic nanobeads. Sci Rep 9:2069. https​://doi. org/10.1038/s4159​8-019-39575​-7 122. Pang KM, Castanotto D, Li H et al (2018) Incorporation of aptamers in the terminal loop of shRNAs yields an effective and novel combinatorial targeting strategy. Nucleic Acids Res 46:e6–e6. https​://doi.org/10.1093/nar/gkx98​0 123. Kaplitt MG, Feigin A, Tang C et al (2007) Safety and tolerability of gene therapy with an adenoassociated virus (AAV) borne GAD gene for Parkinson’s disease: an open label, phase I trial. Lancet 369:2097–2105. https​://doi.org/10.1016/S0140​-6736(07)60982​-9 124. Katada Y, Kobayashi K, Tsubota K, Kurihara T (2019) Evaluation of AAV-DJ vector for retinal gene therapy. PeerJ 7:e6317. https​://doi.org/10.7717/peerj​.6317 125. Yao J, Rotenberg D, Whitfield AE (2019) Delivery of maize mosaic virus to planthopper vectors by microinjection increases infection efficiency and facilitates functional genomics experiments in the vector. J Virol Methods 270:153–162. https​://doi.org/10.1016/j.jviro​met.2019.05.010 126. Forjanic T, Markelc B, Marcan M et  al (2019) Electroporation-induced stress response and its effect on gene electrotransfer efficacy. In  vivo imaging and numerical modeling. IEEE Trans Biomed Eng 66:2671–2683. https​://doi.org/10.1109/TBME.2019.28946​59 127. Schmitt MA, Friedrich O, Gilbert DF (2019) Portoporator©: a portable low-cost electroporation device for gene transfer to cultured cells in biotechnology, biomedical research and education. Biosens Bioelectron 131:95–103. https​://doi.org/10.1016/j.bios.2019.02.024 128. Kasala D, Yoon A-R, Hong J et  al (2016) Evolving lessons on nanomaterial-coated viral vectors for local and systemic gene therapy. Nanomedicine 11:1689–1713. https​://doi.org/10.2217/ nnm-2016-0060 129. Pinyon JL, Klugmann M, Lovell NH, Housley GD (2019) Dual-plasmid bionic array-directed gene electrotransfer in HEK293 cells and cochlear mesenchymal cells probes transgene expression and cell fate. Hum Gene Ther 30:211–224. https​://doi.org/10.1089/hum.2018.062 130. Durymanov M, Reineke J (2018) Non-viral delivery of nucleic acids: insight into mechanisms of overcoming intracellular barriers. Front Pharmacol 9:1–15. https​://doi.org/10.3389/fphar​ .2018.00971​ 131. Hardee C, Arévalo-Soliz L, Hornstein B, Zechiedrich L (2017) Advances in non-viral DNA vectors for gene therapy. Genes (Basel) 8:65. https​://doi.org/10.3390/genes​80200​65 Reprinted from the journal

45

13



Topics in Current Chemistry (2020) 378:13

132. González B, Ruiz-Hernández E, Feito MJ et  al (2011) Covalently bonded dendrimer-maghemite nanosystems: nonviral vectors for in vitro gene magnetofection. J Mater Chem 21:4598. https​://doi. org/10.1039/c0jm0​3526b​ 133. Sohrabijam Z, Saeidifar M, Zamanian A (2017) Enhancement of magnetofection efficiency using chitosan coated superparamagnetic iron oxide nanoparticles and calf thymus DNA. Colloids Surf B Biointerfaces 152:169–175. https​://doi.org/10.1016/j.colsu​rfb.2017.01.028 134. Cen C, Wu J, Zhang Y et al (2019) Improving magnetofection of magnetic polyethylenimine nanoparticles into MG-63 osteoblasts using a novel uniform magnetic field. Nanoscale Res Lett 14:90. https​://doi.org/10.1186/s1167​1-019-2882-5 135. Govindarajan S, Kitaura K, Takafuji M et  al (2013) Gene delivery into human cancer cells by cationic lipid-mediated magnetofection. Int J Pharm 446:87–99. https​://doi.org/10.1016/j.ijpha​ rm.2013.01.055 136. Mu X, Li J, Yan S et al (2018) siRNA delivery with stem cell membrane-coated magnetic nanoparticles for imaging-guided photothermal therapy and gene therapy. ACS Biomater Sci Eng 4:3895– 3905. https​://doi.org/10.1021/acsbi​omate​rials​.8b008​58 137. Namgung R, Singha K, Yu MK et  al (2010) Hybrid superparamagnetic iron oxide nanoparticlebranched polyethylenimine magnetoplexes for gene transfection of vascular endothelial cells. Biomaterials 31:4204–4213. https​://doi.org/10.1016/j.bioma​teria​ls.2010.01.123 138. Kami D, Takeda S, Makino H et  al (2011) Efficient transfection method using deacylated polyethylenimine-coated magnetic nanoparticles. J Artif Organs 14:215–222. https​://doi.org/10.1007/ s1004​7-011-0568-6 139. Wu H-C, Wang T-W, Bohn MC et al (2010) Novel magnetic hydroxyapatite nanoparticles as nonviral vectors for the glial cell line-derived neurotrophic factor gene. Adv Funct Mater 20:67–77. https​://doi.org/10.1002/adfm.20090​1108 140. Prijic S, Prosen L, Cemazar M et al (2012) Surface modified magnetic nanoparticles for immunogene therapy of murine mammary adenocarcinoma. Biomaterials 33:4379–4391. https​://doi. org/10.1016/j.bioma​teria​ls.2012.02.061 141. Veiseh O, Kievit FM, Gunn JW et al (2009) A ligand-mediated nanovector for targeted gene delivery and transfection in cancer cells. Biomaterials 30:649–657. https​://doi.org/10.1016/j.bioma​teria​ ls.2008.10.003 142. Castillo B, Bromberg L, López X et  al (2012) Intracellular delivery of siRNA by polycationic superparamagnetic nanoparticles. J Drug Deliv 2012:1–12. https​://doi.org/10.1155/2012/21894​0 143. Kievit FM, Veiseh O, Bhattarai N et  al (2009) PEI–PEG–chitosan–copolymer-coated iron oxide nanoparticles for safe gene delivery: synthesis, complexation, and transfection. Adv Funct Mater 19:2244–2251. https​://doi.org/10.1002/adfm.20080​1844 144. Mykhaylyk O, Antequera YS, Vlaskou D, Plank C (2007) Generation of magnetic nonviral gene transfer agents and magnetofection in vitro. Nat Protoc 2:2391–2411. https​://doi.org/10.1038/nprot​ .2007.352 145. Gulce-Iz S, Saglam-Metiner P (2019) Current state of the art in DNA vaccine delivery and molecular adjuvants: Bcl-xL anti-apoptotic protein as a molecular adjuvant. In: Immune response activation and immunomodulation. IntechOpen. https​://doi.org/10.5772/intec​hopen​.82203​ 146. Boxus M, Tignon M, Roels S et  al (2007) DNA immunization with plasmids encoding fusion and nucleocapsid proteins of bovine respiratory syncytial virus induces a strong cell-mediated immunity and protects calves against challenge. J Virol 81:6879–6889. https​://doi.org/10.1128/ JVI.00502​-07 147. Al-Deen FN, Ho J, Selomulya C et al (2011) Superparamagnetic nanoparticles for effective delivery of malaria DNA vaccine. Langmuir 27:3703–3712. https​://doi.org/10.1021/la104​479c 148. Garu A, Moku G, Gulla SK, Chaudhuri A (2016) Genetic immunization with in  vivo dendritic cell-targeting liposomal DNA vaccine carrier induces long-lasting antitumor immune response. Mol Ther 24:385–397. https​://doi.org/10.1038/mt.2015.215 149. Tyagi S, Kramer FR (1996) Molecular beacons: probes that fluoresce upon hybridization. Nat Biotechnol 14:303–308. https​://doi.org/10.1038/nbt03​96-303 150. Liu H, Li S, Tian L et al (2010) A novel single nucleotide polymorphisms detection sensors based on magnetic nanoparticles array and dual-color single base extension. J Nanosci Nanotechnol 10:5311–5315. https​://doi.org/10.1166/jnn.2010.2386 151. Lapitan LDS, Xu Y, Guo Y, Zhou D (2019) Combining magnetic nanoparticle capture and polyenzyme nanobead amplification for ultrasensitive detection and discrimination of DNA single nucleotide polymorphisms. Nanoscale 11:1195–1204. https​://doi.org/10.1039/C8NR0​7641C​

13

46

Reprinted from the journal

Topics in Current Chemistry (2020) 378:13 152. Lee M-H, Leu C-C, Lin C-C et  al (2019) Gold-decorated magnetic nanoparticles modified with hairpin-shaped DNA for fluorometric discrimination of single-base mismatch DNA. Microchim Acta 186:80. https​://doi.org/10.1007/s0060​4-018-3192-9 153. Sharma R, Akshath US, Bhatt P, Raghavarao K (2019) Fluorescent aptaswitch for chloramphenicol detection—quantification enabled by immobilization of aptamer. Sens Actuators B Chem 290:110– 117. https​://doi.org/10.1016/j.snb.2019.03.093 154. Xuhong Y, Sinong Z, Jianping L et al (2019) A PCR-lateral flow assay system based on gold magnetic nanoparticles for CYP2C19 genotyping and its clinical applications. Artif Cells Nanomed Biotechnol 47:636–643. https​://doi.org/10.1080/21691​401.2019.15758​41 155. Cheng H, Liu J, Ma W et al (2018) Low background cascade signal amplification electrochemical sensing platform for tumor-related mRNA quantification by target-activated hybridization chain reaction and electroactive cargo release. Anal Chem 90:12544–12552. https​://doi.org/10.1021/acs. analc​hem.8b024​70 156. Shan Y, Zhang Y, Kang W et  al (2019) Quantitative and selective DNA detection with portable personal glucose meter using loop-based DNA competitive hybridization strategy. Sens Actuators B Chem 282:197–203. https​://doi.org/10.1016/j.snb.2018.11.062 157. Tian B, Qiu Z, Ma J et al (2018) On-particle rolling circle amplification-based core–satellite magnetic superstructures for microRNA detection. ACS Appl Mater Interfaces 10(3):2957–2964. https​ ://doi.org/10.1021/acsam​i.7b162​93 158. Li W, Jiang W, Dai S, Wang L (2016) Multiplexed detection of cytokines based on dual bar-code strategy and single-molecule counting. Anal Chem 88:1578–1584. https​://doi.org/10.1021/acs. analc​hem.5b030​43 159. Xu Y, Huo B, Li C et al (2019) Ultrasensitive detection of staphylococcal enterotoxin B in foodstuff through dual signal amplification by bio-barcode and real-time PCR. Food Chem 283:338– 344. https​://doi.org/10.1016/j.foodc​hem.2018.12.128 160. Jiang P, Haji C, Ye X et al (2019) A novel inductive coupled plasma mass spectrometry gene detection method based on AuNPs and bio-barcode signal amplification. Nanosci Nanotechnol Lett 11:638–644. https​://doi.org/10.1166/nnl.2019.2933

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Affiliations José Raúl Sosa‑Acosta1,2 · Claudia Iriarte‑Mesa1 · Greter A. Ortega1,3 · Alicia M. Díaz‑García1  1

Laboratory of Bioinorganic (LBI), Department of Inorganic and General Chemistry, Faculty of Chemistry, University of Havana, Havana, Cuba

2

Faculty of Chemistry and Pharmacy, Pontifical Catholic University of Chile, Santiago, Chile

3

Legaria Unit, Center for Applied Science and Advanced Technology of IPN, Mexico City, Mexico



Reprinted from the journal

47

13

Topics in Current Chemistry (2020) 378:40 https://doi.org/10.1007/s41061-020-00302-w REVIEW

Magnetic Nanoparticles as MRI Contrast Agents Ashish Avasthi1 · Carlos Caro1 · Esther Pozo‑Torres2 · Manuel Pernia Leal2 · María Luisa García‑Martín1,3 Received: 31 December 2019 / Accepted: 18 March 2020 / Published online: 7 May 2020 © Springer Nature Switzerland AG 2020

Abstract Iron oxide nanoparticles (IONPs) have emerged as a promising alternative to conventional contrast agents (CAs) for magnetic resonance imaging (MRI). They have been extensively investigated as CAs due to their high biocompatibility and excellent magnetic properties. Furthermore, the ease of functionalization of their surfaces with different types of ligands (antibodies, peptides, sugars, etc.) opens up the possibility of carrying out molecular MRI. Thus, IONPs functionalized with epithelial growth factor receptor antibodies, short peptides, like RGD, or aptamers, among others, have been proposed for the diagnosis of various types of cancer, including breast, stomach, colon, kidney, liver or brain cancer. In addition to cancer diagnosis, different types of IONPs have been developed for other applications, such as the detection of brain inflammation or the early diagnosis of thrombosis. This review addresses key aspects in the development of IONPs for MRI applications, namely, synthesis of the inorganic core, functionalization processes to make IONPs biocompatible and also to target them to specific tissues or cells, and finally in vivo studies in animal models, with special emphasis on tumor models. Keywords  Magnetic nanoparticles · Iron oxide nanoparticles · Magnetic resonance imaging · Cancer · Diagnosis

Chapter 3 was originally published as Avasthi, A., Caro, C., Pozo‑Torres, E., Leal, M. P. & García‑Martín M. L. Topics in Current Chemistry (2020) 378: 40. https://doi.org/10.1007/s41061-020-00302-w. * Manuel Pernia Leal [email protected] * María Luisa García‑Martín [email protected] 1

BIONAND ‑ Centro Andaluz de Nanomedicina y Biotecnología, Junta de AndalucíaUniversidad de Málaga, C/Severo Ochoa, 35, 29590 Málaga, Spain

2

Departamento de Química Orgánica y Farmacéutica, Facultad de Farmacia, Universidad de Sevilla, 41012 Seville, Spain

3

Networking Research Center on Bioengineering, Biomaterials and Nanomedicine, CIBER-BBN, Málaga, Spain



Reprinted from the journal

49

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:40

1 Introduction Magnetic resonance imaging (MRI) is one of the main in vivo imaging modalities, along with positron emission tomography (PET), computed tomography (CT) and ultrasound imaging. MRI is the most versatile of all of these, being able to provide both anatomical and functional information with excellent image quality, and, most importantly, using non-ionizing radiation, which allows longitudinal studies to be performed without the risk of side effects. The MRI signal comes from the radiofrequency signal of protons magnetized by an external magnetic field. These protons originate mainly from water molecules. The application of radiofrequency pulses is used to excite the magnetization, and magnetic field gradients are used to provide spatial localization. Contrast in MRI reflects differences in signal intensity, which depends on the concentration of water molecules within the tissue, the relaxation times, ­T1 and ­T2, of the water protons and the mobility of the water molecules (diffusion, flow) [1]. Additionally, image contrast can be further enhanced using contrast agents (CAs), with Gd-chelates being used most commonly in clinical practice. However, CAs lack specificity and have recently been related to toxicity issues caused by the unexpected release of free Gd. Magnetic nanoparticles have emerged as a promising alterative with improved properties in terms of specificity and biocompatibility. Over the past two decades, many studies have aimed at the development of new magnetic nanomaterials that can serve to improve the diagnosis and treatment of many different diseases. Among these nanomaterials, iron oxide nanoparticles (IONPs) have been investigated most extensively as CAs for MRI due to their magnetic properties, that is, the superparamagnetism that leads to very high relaxivity, their high biocompatibility, since they can be incorporated into iron metabolism, and also the easy functionalization of their surfaces with target molecules for molecular imaging purposes [2]. The first step in the development of IONPs is synthesis of the magnetic core, for which many different methods have been proposed, all aiming at strict control of the size, shape and magnetic properties, so that the synthesis process can be performed under highly reproducible conditions, which is one of the essential requirements for the potential clinical translation of these new nanomaterials [3]. Functionalization of magnetic nanoparticles is then needed to make them soluble in aqueous media and to provide them with stability and biocompatibility [4]. Further functionalization may include the addition of different molecules to target specific tissues or cells [5]. The most relevant functionalization strategies will be discussed in detail in this review. Finally, the in vivo characterization of IONPs is the most critical aspect in the development of IONPs for biomedical applications. Although many new nanomaterials show excellent in vitro properties, most of them fail when tested in vivo. Thus, around 6500 studies (PubMed database) on magnetic nanoparticles have been published since 2010, in which IONPs often appear as promising new CAs for MRI. However, up to now, extremely low clinical translation has been achieved [6]. Therefore, comprehensive studies with appropriate in  vivo experimental models are of paramount importance for the successful development and eventual clinical translation of these nanomaterials.

13

50

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

In this review, we describe the recent advances in regard to the synthesis, functionalization and in  vivo applications of IONPs as MRI CAs for the diagnosis of several pathologies, with special emphasis on cancer diagnosis.

2 Methods for the Synthesis of IONPs Over the past few decades, various procedures to synthesize IONPs have come to fruition. The ultimate goal of these procedures is to gain complete control over the properties of IONPs, such as size, shape, saturation magnetization, etc. However, this has not yet been achieved completely. The main hindrance behind this failure is the inability to fully determine the science behind the processes and their mutual interactions, but it is not so distant in the future that we will be successful. Figure 1 shows different methods to synthesize IONPs, which are described in detail below, along with their pros and cons. 2.1 Coprecipitation Coprecipitation is the method most commonly used for the synthesis of IONPs due to its facile nature. Massart [7] pioneered the existing scientific knowledge established by Le Fort [8] and Elmore [9] regarding the synthesis of magnetic colloids, and stressed the importance of the stoichiometric ratio between Fe(II):Fe(III) being 1:2. The synthesis process described by Massart requires the addition of alkaline medium (pH ~ 11, slowly or rapidly) into the iron salts solution at room temperature or at elevated temperature. This mixture requires an inert atmosphere to prevent nanoparticles from oxidizing. It was later established that the synthesis of particles follows the LaMer’s model of nucleation and growth [10]  (Fig.  2). The synthesis process has been described to occur in two steps, as shown below [11–14]

Fe2+ + 2Fe3+ + 8OH− ⇆Fe(OH)2 + 2Fe(OH)3 → Fe3 O4 ↓ + 4H2 O. However, Lagrow et  al. [15] recently challenged this mechanism of synthesis. They claimed that while increasing the pH via sodium carbonate, two intermediate phases are formed, one poorly crystalline ferrihydrite and another crystalline iron hydroxide carbonate. This ferrihydrite eventually grows into iron oxide at the cost of iron hydroxy carbonate. Even though Lagrow’s proposed mechanism seems to answer a few loopholes undescribed by Massart, improving the homogeneity and reproducibility of the nanoparticles, it fails to ascertain if the same mechanism is followed when ammonia or ammonium hydroxide is used. Irrespective of the mechanism followed, nucleation is judged as the sizedetermining step and is exploited to modulate the size of particles [14–16]. The nature of particles depends on various other factors, such as the type of salts used (e.g. chlorides, sulfates, nitrates, perchlorates, etc.), the ­Fe2+ and ­Fe3+ ratio, pH and the ionic strength of the media, along with the reaction environment [17–30]. Jiang et al. [24] showed that the particle size distribution is narrowed Reprinted from the journal

51

13



Topics in Current Chemistry (2020) 378:40

Fig. 1  Methods used in the synthesis of iron oxide nanoparticles (IONPs)

if the homogeneity of pH within the solution is improved by adding urea to the reaction mixture. There are also reports suggesting that particle size decreases with increasing pH [17]. A similar trend is observed between particle stability and iron concentration, but substantial studies are lacking to support this observation [23]. Particles with different morphologies, such as nanodots, ellipsoid,

13

52

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



Fig. 2  LaMer’s model depicting the nucleation and growth process of the nanoparticles. Adapted with permission from [10]. Copyright (1950) American Chemical Society

spherical, clusters or necklace like, can be synthesized by varying their aging conditions [25–27]. Itoh et al. [26] synthesized ellipsoidal and spherical hematite nanoparticles by aging them in phosphate ions and nitriloacetic acid (NTA), respectively. The relationship between shape/size and the electrostatic surface density of particles is linked to the interfacial tension between the oxide and the solution, which causes a decrease in the surface energy, thus modulating shape and size [28]. If a modern method like ultrasonication is used with coprecipitation, it can yield narrowly distributed particles, as shown by Bui et  al. [31], who compared their modified version of the coprecipitation method (using ultrasonication instead of stirring) to the solvothermal method, and found the former to yield more homogeneous and small sized nanoparticles. However, the comparison between their method and the conventional coprecipitation method (with stirring) is missing. The major advantages of the coprecipitation method are its time saving facile nature, with no requirement of high temperature or pressure, and the production of particles with high yield and easily scalable to large quantities. However, the particles synthesized with this method generally lack homogeneity and form single and also multicore nanoparticles. Particles thus synthesized also tend to form aggregates, which leads to an undesired assortment of blocking temperatures. Another disadvantage of this method is that the pH of the resultant solution is too high, thus requiring neutralization before they can be used for biological applications.

Reprinted from the journal

53

13



Topics in Current Chemistry (2020) 378:40

2.2 Thermal Decomposition In this method of synthesis, high temperatures are exploited to break down the precursor to yield nuclei as well as their further growth into nanoparticles (Fig. 3). It started as a way to ease the study of properties of systems with narrow size distribution [32]. Smith and Wychlk were among the first researchers who utilized this method to synthesize colloidal dispersions of iron using iron pentacarbonyl [Fe(CO)5] as a precursor, along with different solvents and the addition of different polymers. They concluded that the polymers added during the reaction not only coated the dispersions forming stable particles, but also acted as catalysts for the decomposition [33, 34]. They suggested that the decomposition takes place at 140–160 ℃ in the presence of butadiene polymers while gathering support from the mechanistic studies conducted by Bergman and coworkers [35]. Later, their hypothesis was verified experimentally, showing the presence of an intermediate carbonyl complex formed after decomposition of Fe(CO)5 [36]. The reaction takes place in two main steps: nucleation and growth. This separation of stages can be used advantageously to alter the size and shape of nanoparticles as demonstrated by Hyeon et al. [37] and Jana et al. [38]. They used iron oleate as precursor and proposed that nucleation starts at 200–240 ℃, initiated by dissociation of one of the three oleates available in one molecule of iron oleate [Fe-(oleate)3], while the growth begins at 300 ℃ with the subsequent dissociation of the remaining two oleates. The complete mechanism of the reaction is not fully understood even though it has been widely studied, both experimentally and computationally [39–41]. Nonetheless, these studies led to the discovery of “polyiron oxo clusters” species as the actual precursor for the formation of nanoparticles, as initially suggested by Wells [36]. More recent studies have reported the synthesis of a new precursor by synthesizing an intermediate between Fe(CO)x and oleylamine (OLA), and achieved controllable size of 2.3–10 nm [42]. To date, different precursors have been reported in the literature: iron acetylacetonate [Fe(acac)3] [43], iron cupferron [Fe(cup)] [44], iron chloride (­FeCl3) [45], iron pentacarbonyl [Fe(CO)5] [46], along with different iron complexes such as iron oleate [45], iron stearate [38] and iron eruciate [47]. Depending on the process involved and the size required, it becomes important to select the right

Fig. 3  Different stages during the synthesis of IONPs in the thermal decomposition method. Adapted and modified with permission from [41]. Copyright (2013) American Chemical Society

13

54

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



precursor as the reaction proceeds differently depending on the way the precursor is broken down [48]. There are several other factors that could affect the size and morphology of particles, such as temperature, nature of the solvent, reactants ratio, reflux time, and seed concentration [45, 49, 50]. Thus, Hyeon found that the heating rate of the reaction, along with the boiling point  of the solvent used, is also a crucial factor to adjust the size of the nanoparticles [45], and Pellegrino’s group concluded that there is an inverse relationship between the size of the nanoparticles and the heating rate [49]. However, controversy still exists regarding the role of the temperature ramp in the synthesis of IONPs, and, therefore, comprehensive and deeper studies are still needed to properly elucidate the mechanism involved. Kovalenko et al. [51] showed the importance of surfactants, not only to prevent aggregation, but also to modulate shape and size. They displayed the use of fatty acids, such as oleic acid (OA) or salts of OA, to synthesize spheres and cubic nanoparticles, respectively. Later, several groups have tried to shed light on the role of OA as well as other fatty acids regarding the size and shape of IONPs, but up to now, a fully elucidated theory is still lacking [52–59]. Quality of particles can be further improved by the controlled addition of water and oxygen in the inert environment to decrease crystal defects, and improve magnetic properties and homogeneity [60, 61]. In summary, thermal decomposition, albeit a bit complex and time-consuming, yields very homogenous and monodisperse nanoparticles, making it one of the most used methods to synthesize nanoparticles for biological applications. The shape and size of nanoparticles can be controlled by tuning the parameters described above. Major drawbacks of this method include the inability to properly scale up and the lack of dispersibility of the particles in aqueous solvents, although this can be remedied by surface modifications in situ, as described by Li et al. [56, 62], or using post preparative methods, as explained in greater detail in later sections of this review. 2.3 Hydrothermal and Solvothermal Synthesis In this method, the hydrolysis and oxidation (or neutralization) reaction takes place in a reactor or autoclave at high temperature and pressure. Depending on the reaction solvent, it is either referred to as hydrothermal (if the solvent is water) or solvothermal (any other solvent or combination). Both reactions follow the aforementioned model of nucleation and growth [63, 64]. There have been several reports [4, 37, 65–68] on the use of this method to synthesize magnetic nanoparticles as well as its comparison with other methods [69]. The reaction parameters, such as temperature, reactor size, time, concentration of the reactants, and the nature of the solvent and capping agents, affect the size, shape and other properties of the final product. Out of all these parameters, the effect of the solvent has been studied the most [70, 71], closely followed by that of the surfactant [72, 73]. The particles show a preferential surface binding towards the carboxylate from the OA rather than the amine from the oleylamine [72], which very likely is the case for every method described in this article, although it still needs verification.

Reprinted from the journal

55

13



Topics in Current Chemistry (2020) 378:40

This preferential binding was recently used by Brewster et al. [73] to present a new way to control the particle size and crystal phase. They varied the carbon chain length in the iron carboxylate, which was used as the precursor, and showcased the effect of two different ligands, amine and carboxylic acid, which were added to the reaction [73]. They demonstrated that the size of the particles decreased as the carboxylate chain length increased in the presence of amine ligands, while no definite trend was observed when varying the carboxylate free ligands. The hydrothermal/solvothermal method has also been used to synthesize other ferrites [74]. Kim et  al. [75] recently demonstrated a gram scale yield of magnetite nanoclusters by modifying the procedure and utilizing trisodium dihydrate, but, to the best of our knowledge, this is the only report for large scale synthesis using this method. To further exploit the particles thus formed for biological applications, surface coating becomes necessary, as will be discussed in detail in the subsequent section. Polymers such as polyvinylpyrrolidone (PVP), polyacrylic acid (PAA) and polyethanolimine (PEI), have been shown to improve the magnetic properties when used in the synthesis of monodispersed clusters [76]. Recently, Köçkar et  al. [77] explained a way to get in-situ capping of IONPs with tartaric acid/ascorbic acid/ mixture of two, which led to the synthesis of uniform, un-agglomerated, biocompatible particles of less than 8 nm with good saturation magnetization. The hydrothermal/solvothermal method is, therefore, an ideal method for the synthesis of iron oxide nanoparticles, mainly nanoclusters. However, the main disadvantage of this method is that, due to the lack of stirring inside the autoclave, monodispersity, as well as scalability, can sometimes be hindered. 2.4 Polyol Method This method is an iteration of the solvothermal method, with polyols being used as solvents to synthesize nanoparticles by dissolving the precursor, solubilizing in the diol at high temperatures, and eventually leading to the formation of metal nuclei and particles. Following previous works pertaining to synthesis of metallic powders [78–84], Caruntu et  al. described this method to synthesize nanocrystalline metal oxide nanoparticles by synthesizing magnetite nanoparticles [85]. They explained the mechanism stating that reduction starts from the liquid state rather than the solid, and the nanoparticles are formed in two steps: hydroxides are formed first and then metal centers are chelated. Heterogeneous nucleation performs better than homogeneous nucleation as it has been studied to provide a better separation between nucleation and growth, thus giving better control over the size, shape and crystallinity [78]. Polyols play multiple roles, acting as reducing agent, stabilizer and solvent [86], modulating the process to yield large and small clusters [87], nanoparticles [88] or single-core/multicore nanoparticles [89]. Different polyols have been exploited for the synthesis of iron oxide nanoparticles, such as diethylene glycol, giving 3  nm particles [90], or triethylene glycol, giving 10  nm particles [91]. However, Cai et al. [92] reported that only triethylene glycol gives nonaggregated nanoparticles. To our knowledge, there are no reports on the use of tetra or penta ethylene glycol, which could have ameliorated the agglomeration problem

13

56

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

even more, if the trend described holds to be true. Other parameters that have been identified to modulate the size, shape, crystallinity and saturation magnetization are temperature, time, precursor concentration, and surfactant. The role of water was studied by Hemery et al. [93], when its importance was revealed by inability of the anhydrous iron chloride to produce magnetic particles [93]. The impact of stoichiometry in polyol synthesis has been studied by Wetegrove et al. [94], showing that the increase in F ­ e3+ concentration forms larger crystallites and the increase in F ­ e2+ content promotes nucleation [94]. As stated above, hydrophilicity is important, which is generally lacking in particles synthesized using the polyol method. However, there are reports of the synthesis of hydrophilic nanoparticles using this method [88, 95] but their limitations include the lack of a surface functionality for bioconjugation. This problem has further been remedied by the use of polyamines [96], polyimine with polyol [97], polyamine with polyol [98] and PAA [99]. The research done by Babić-Stojić et al. [100], wherein they esterified 3 nm IONPs in situ, implied the importance of the surface layer in the properties of nanoparticles. The morphology of the particles is of equal importance as size in in vivo applications and has been shown to be altered by the addition of halide ions [101]. There have also been advancements in solvents, such as the thermostable ionic solvent [P6,6,6,14][Tf2N], which has been shown to be capable of synthesizing quasi spherical magnetite nanoparticles of around 14 nm [102]. In conclusion, the method described herein has the advantage of being environment friendly, scalable, and good for synthesizing both single and multicore particles. However, it has the drawback that the particles thus formed lack homogeneity. 2.5 Sol–Gel Method This is a two-step chemical method, with the first step being the synthesis of the sol (particles in a solution) via hydroxylation of the precursors, and the second step, the formation of a gel by condensation and polymerization. Eventually, heat treatments are used to achieve a proper crystalline state. Costa et  al. [19] were among the first to synthesize magnetic nanoparticles using this method, but they failed to identify the correct mechanism. Subsequently, the work of Portugal et  al. [103], made the mechanism a bit clearer upon finding signatures of iron hydroxide, but the exact mechanism is still unknown. Like in the polyol method, the solvent is shown to affect the ferrite grain as well, but changes in grain size have been attributed to a different growth model with two different solvents [104]. Water concentration is also shown to improve hardness and structural defects [105]. Size and shape are also affected by other parameters such as solvent ratio, time, pH, stirring, gelating agent and, temperature. Liu et al. [106] used different calcination temperatures to synthesize different phases of IONPs, and this transformation has been attributed to two separate mechanisms, crystal regrowth and chemisorption, depending on the temperature. Akbar et al. claimed to have synthesized three different phases of iron oxide (α-Fe2O3, γ-Fe2O3, and F ­ e3O4) simply by varying the precursor to solvent ratio, thus suggesting the importance of that ratio [107]. The

Reprinted from the journal

57

13



Topics in Current Chemistry (2020) 378:40

particles were shown to possess higher saturation magnetization. They also observed differences in hematite particle size and morphology when using different precursors, with iron acetate giving rise to smaller spherical particles, while iron nitrate led to larger, quasi cubic particles. These differences were due to the water content as well as the presence of nitrate and carboxylate in the precursors [108]. More recently, Hu et al. [109] reported a new explosion-assisted sol–gel method in which they used ferric nitrate as precursor and citric acid as chelating agent to form a gel. The gel was then homogenized and heated with picric acid to attain highly pure, well dispersed and crystallized magnetite nanoparticles ranging from 3 to 20  nm. The synthesis was proposed to be resulting from the combined action of the complexing of citric acid with metal ions, and the explosion, thus explaining the important role of citric acid, not only as a carbon source, but also to allow the combustion and reduction of the dried gel simultaneously. The chemistry of the sol–gel method is vast, with the involvement of different precursors, gelators as well as chelators, but it is beyond the scope of this review. It is, however, nicely explained by Danks et al. [110]. This method is more recommended for synthesizing thin films [111] and nanocomposites [112, 113] since it can form thin films in just 2 min if the heating source is changed to microwaves, and pure phases can be formed by using high microwave power (600–800 W). 2.6 Microemulsion Method The microemulsion method is a form of coprecipitation performed in a confined space such as micelles. It generally involves two immiscible liquids with surfactants forming the interfacial layer [114], and is classified as either the water-in-oil method or oil-in-water method [115]. Inouye et al. [116] were the first to report the synthesis of magnetic particles using this method, exploiting the faster oxidation of ferrous ions in micelles. In water-in-oil microemulsion, a hydrophobic phase is used with aqueous droplets separated by a surfactant [117]. The most common surfactants used are PVP and cetyltrimethylammonium bromide (CTAB). In this method, particles generally collide and coalesce, and break again, leading to the growth of particles, the particle size being determined by the size of the droplets. In a final step, particles are centrifuged and lyophilized to get pure nanoparticles [118–120]. Many articles have been published on the use of this method to synthesize iron oxide nanoparticles [121–124]. Although surfactant concentration is not shown to affect the size, precursor concentration and temperature are important influencers, together with pH [125] and the choice of surfactant [126, 127]. Recently, Singh et al. [128] showed the importance of ionic concentration and temperature on the morphology, size and crystallinity by claiming that, in order to obtain monophasic particles, ­[Fe2+] and ­[Fe3+] should be ≤ 0.09  M and ≤ 0.184  M, respectively, with a temperature range of 65–72 ℃. They also observed changes in the morphology of the particles, from cubes to pentagons to spheres, when increasing the concentration of the surfactant (CTAB) between 0.01 and 0.1  M, but they did not describe

13

58

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

the mechanism, or explain why the shape of the CTAB nanodroplets changes upon varying concentration. Nor did they explain why particle size changed with concentration [128]. Bonachhi et al. [129] achieved ultra-small magnetic nanoparticles by using γ-cyclodextrin by hydrolyzing F ­ e2+ ions in aqueous solution, while Lee et al. [130] varied the ratios of the precursor and solvent from 3.6 to 8.1, and achieved 2 to 10 nm magnetite particles. Vidal et al. showed the importance of oleylamine as surfactant to prevent aggregation [131], while Pileni et al. explained the importance of using functionalized surfactants and pH to improve the crystallinity and morphology of the nanoparticles [132]. Following a similar approach, Han et al. used a nonionic surfactant, ­C16E15, to synthesize nanoparticles with high saturation magnetization (74.8  emu/g) [133]. It is worth mentioning that if the surfactant described in this method is replaced by a phospholipidic molecule to form particles within liposomes, they are termed magnetoliposomes, which show significantly higher blood half-life [134–136]. However, if the particles are formed within the aqueous compartment, they are known as magnetovesicles. These special particles can be synthesized using film hydration and extrusion [137], sonication [66], phase evaporation [138] and nanoreactor [139], and are very promising for biomedical applications. Recently, even metallosurfactants have been used as precursors to synthesize particles of around 3 nm [140]. This method has also been utilized in exchanging the capping of iron oxide nanoparticles to improve solubility [141–143]. Similarly, oil-in-water has a hydrophilic solution with oil droplets used as a reactor. Recently, spinel ferrites have been shown to be synthesized using this method, with metal ethylhexanoates as precursors and a pseudo ternary solvent system, which includes oil, surfactant and water in the ratios of 20:20:60 [144]. The oil in water method has also been used as a strategy to cap nanoparticles [145]. The microemulsion method has several advantages, such as providing a narrow range of particles with relative ease, good morphology and without the need for high temperatures. But it also has disadvantages, including scalability, the toxicity of some surfactants, the amount of surfactant used, as well as the need for ligand exchange. 2.7 Aerosol Method This is also a chemical method, which leads to high production of particles. This method can be subdivided in two categories. The first is spray pyrolysis, in which precursor salts are sprayed into the reactors, where they are condensed and solvent is evaporated, which in turn also means that the size of the particles depends on the droplets [146]. Serna’s group [147] were among the first to synthesize ­Fe2O3 nanoparticles using this method. Their study claimed that if small size is the most important feature for the application, iron acetylacetonate should be used because of its exothermic decomposition reaction; however, if crystallinity is to be considered, then iron chloride is favored due to solvent elimination at higher temperature. This leaves other precursor benefits open for exploration. The importance of intraparticle reactions in controlling the size of particles was established later, along with the solvent, rate Reprinted from the journal

59

13



Topics in Current Chemistry (2020) 378:40

of evaporation, time spent in the reactor, and temperature. These studies concluded that the heating time and temperature, along with the type of evaporation or reaction taking place during the drying stage, will conform the particle structure as hollow, dense, foam-like, etc. [148, 149]. Zheng et al. [150] recently reported that chloride ions prevent phase transition from γ-Fe2O3 to α-Fe2O3 at higher temperatures, leading to higher magnetization, which highlights the importance of chloride ions in the reaction. Das et al. proposed a new strategy to decrease size with high crystallinity by adding ethanol to the ultrasonic pyrolysis [151]. It was explained that the faster evaporation rate of ethanol compared to water, as well as a decrease in surface tension of the water–ethanol solution, led to the formation of smaller droplets and eventually smaller particles. Since the rate of evaporation of the solvent has been stressed and linked to particle size, it might be interesting to see how methanol, or any other solvent with a boiling point lower than that of ethanol, affects the size and crystallinity of particles. The second category is Laser pyrolysis, a gas phase method that utilizes the heat generated by a laser to heat the precursors and the flow of a gas or a mixture of gases to produce nanoparticles. The sizes of the particles can be controlled by modulating the power of the laser since a direct relationship exists between the two [152, 153]. Zhao et al. [154] were the first to improve on the TEA laser using a cw C ­ O2 laser, which yielded particles with higher purity. There have also been reports on use of this method to synthesize hybrid silica-iron oxide composites [155]. Laser pyrolysis has a new iteration, flame spray pyrolysis (FSP), which uses a flame to heat the precursor [156]; the size of the nanoparticles can be controlled by varying the flame length or the oxidant flow rate, and the precursor/fuel composition. Lower flow rate of the oxidant leads to reduced flame length, with higher temperatures thus forming smaller particles and vice versa [157]. The main advantage of this method is that it helps in achieving very high homogeneity and monodispersity irrespective of the complexity of particles, including hybrid silica-iron oxide composites [155]. 2.8 Sonochemical Method This method utilizes acoustic cavitation, which means the formation, growth and collapse of bubbles generated by ultrasound, to synthesize nanoparticles. Instead of using high temperature or pressure directly, this method creates them indirectly by using bubbles or cavities formed in the liquid by the acoustic waves. Further oscillation of such waves helps them gather and store ultrasonic energy, creating a hot spot (~ 5000  K) and leading to the synthesis of particles of different shapes and sizes. This method works for both volatile and non-volatile solvents [158–160]. The reaction medium was already considered the most important factor in controlling the properties of nanoparticles by Suslick et al. [160] when they proposed the method, since the bubbles formed will depend on the vapor pressure of the media. The nature of the particles can also be altered by changing the ultrasonic frequencies based on the inverse relationship between oxidation of F ­ e2+ to ­Fe3+ and ultrasonic frequencies

13

60

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

[161]. The synthesis of particles in the presence of different ligands has also been performed, giving rise to particles between 5 and 16 nm [162]. Vijayakumar et al. [163] used a similar route to synthesize IONPs. They proposed a mechanism stating that ultrasonic waves produce the vaporization of water and further pyrolyzation into H and OH radicals due to prolonged temperature and pressure, which leads to the formation of hydrogen ­(H2) and hydrogen peroxide ­(H2O2) from the reaction between ­H2 and hydroxyl radicals, respectively. Meanwhile, the same energy also breaks down iron acetate into Fe(II) ions. These Fe(II) ions are later oxidized to Fe(III) using H ­ 2O2 as oxidant and forming F ­ e3O4 by using OH radicals [163]. There are several studies showing the effect of surfactants on the particles. Mukh-Qasim et  al. [164] used SDS as stabilizer to get around 8.5  nm amorphous but water dispersible ­Fe3O4 particles, while Rahamwati et al. used iron sands along with different concentrations of PEG-6000. This latter group showed that, as PEG concentrations increased, the crystallite size of the particles increased [165]. They also showed that the morphology of the particles shifted from flower-like to cubes to spheres with increasing PEG concentrations. Kim et  al. [166] synthesized OA-capped IONPs, which form a ferrofluid when dispersed in chitosan, with a hydrodynamic diameter of 65  nm, thus being potential MRI CAs. This method can also be used to synthesize composite nanoparticles [167] or other ferrites [168, 169].This method has also been used for surface functionalization in very short time [170]. The main advantage of this method is its accelerated nature to produce nanoparticles with good yield, but it falls short when it comes to phase homogeneity. 2.9 Microwave Synthesis This is a modern-day hydrothermal method of synthesizing nanoparticles and one of the most used in recent days due to its much-improved kinetics of crystallization. It requires as low as 10 s and yields small and monodisperse particles due to homogenous heating [171]. Palchik et al. were among the first to use this method in a domestic microwave oven and suggested that the synthesis of particles was happening due to thermal breakdown of Fe(CO)5, which in turn was taking place due to heating of chlorobenzene, since Fe(CO)5 is a microwave resistant compound [172]. This indirectly marks the importance of the solvent. On the other hand, Liu et al. demonstrated the importance of water in maintaining a stable heating environment, along with the role of stoichiometry [173]. Another important parameter that have been studied extensively is the nature of the surfactant, with studies reporting the use of different concentrations of OA [174], amino acids [175], polyethylene glycol (PEG) [176], and different ratios of OA and oleylamine (OLA) [177]. OA is shown to increase saturation magnetization with increasing concentration, with no definite trend in size. However, concentrations beyond 0.35 mmol/dm3 led to agglomeration and the product became difficult to isolate. Recently, amino acids such as glycine have been shown to reduce the crystallite size of IONPs, opening the path to explore other amino acids [175]. The presence of PEG in the reaction has also been shown to lead to smaller IONPs,

Reprinted from the journal

61

13

Topics in Current Chemistry (2020) 378:40

as compared to the reaction in its absence. When the reaction is performed in the presence of PEG, it tends to favor the formation of magnetite instead of maghemite. This happens due to PEG being sacrificial in nature and thus preventing oxidation. High microwave power and low synthesis time also favors the formation of maghemite [176]. Other studies have shown that the presence of OA during the synthesis, along with OLA, reduces aggregation among particles [177]. Temperature has also been shown to transform phases in IONPs [178]. Blanco-Andujar et  al. proposed a facile method to synthesize citric acid coated IONPs and potentially scale them up [179]. The importance of aging temperature on crystallinity can be seen when ­Fe2O3 nanocubes are synthesized by decomposing iron oleate in a microwave and aging it in an autoclave at 180 ℃ for different time intervals [180]. Particles aged for 20 h showed cubic shape and higher saturation magnetization. Hu et al. [181] argued that the precursor is the most important parameter by synthesizing three phases of iron oxide, hematite, magnetite and maghemite, using ­FeCl3 alone or in combination with ­FeCl2. Literature suggests that the morphology and composition of the particles can also be controlled using this method. Different morphologies, such as lamellar sheets [182], octahedrons [182] and hexagonal plates [183] are synthesized by slight changes in salts. Cu-doped IONPs with good colloidal stability are obtained in 10 min [184] using the microwave method. In fact, even using a domestic microwave, sizes of 8–10 nm can be easily achieved [185]. This method has been shown to be better than hydrothermal [186] or thermal decomposition [187] in terms of size, crystallinity and saturation magnetization. However, particles thus synthesized display lower surface reactivity than those synthesized using the thermal decomposition method, although with more ease of stabilization. The versatility of this method is acknowledged by its association with different methods: coprecipitation [179], thermal decomposition, [177] polyol [188] and sol–gel methods [189]. The particle size can be varied by modulating the power and hence the temperature, the time spent in the reactor, the cooling rate, etc. This method has become more popular recently due to its multiple advantages. 2.10 Biosynthesis This is an eco-friendly method as most of the constituents needed are available from nature directly or indirectly. It generally involves the use of microbes [190] or plant extracts [191] to synthesize nanoparticles. Lovely et  al. [192] were the first to use a microbe, GS-15, to form magnetite nanoparticles. Thereafter, many different magnetic bacterial strains were found and studied in order to produce IONPs [193–197].These nanoparticles are formed by the reduction/hydrolyzing capabilities of these biological entities. However, when a bacterium is used, its nature as well as its incubation time becomes an important parameter since it allows changes in size and morphology [198, 199]. Even fungi such as Fusarium oxysporum and Verticillium sp., have been shown to possess hydrolyzing capabilities to form different sizes and shapes of nanoparticles [200]. Viruses such as tobacco mosaic virus (TMV) have also been used as templates to synthesize nanotubes [201]. Iron oxides formed by microbial reduction have been

13

62

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

shown to lead to phase transformation with better crystallinity, although decreasing their reducibility [202].The scalability issue has also been answered by using a 30-l reactor, although there is only one report of such nature [203]. Plants or plants extracts have been used for the synthesis of nanoparticles [204]. Most recently, IONPs have been synthesized using figs, Ficuscarica and Plantago major extracts, which, apart from reducing precursors, also cap and stabilize the particles. These reactions have been concluded to take place due to the presence of phenols, and normally lead to sizes ranging from 2 to 50 nm [205, 206]. The main advantages of this method are that it is energy saving and non-toxic. Also, there is an unlimited supply of reducing agents, making it economically viable. On the other hand, its major disadvantage is unpredictability regarding the nature of the particles, with less control over the shape and size, along with uncertainty of yielding monodisperse particles when scaled up. 2.11 Other Methods Several different methods for the synthesis of IONPs have not been described above due to a dearth of information in the literature. Alvarez et al. [207] developed a novel flow injection synthesis (FIS) method to fabricate magnetite nanoparticles in a capillary reactor, and produced homogenous particles of 2–7 nm with high reproducibility. There have been reports of the use of metal rods as anodes and electrochemical deposition in the presence of surfactants to yield 3–8 nm particles [208–210]. Chemical vapor deposition (CVD) [211, 212] has been used to fabricate thin films and morphologycontrolled nanoparticles. Other methods, such as synthesis in a reactor [213], the solution combustion method [214], and the use of microfluidic channels on a chip [215, 216], have also been introduced. All the methods described above have their own pros and cons, and the choice of one or the other depends on the application for which the nanoparticles are being developed. Thus, for nanoparticles to be used as MRI CAs, the most suitable methods appear to be the thermal decomposition or microwave methods, since they provide a very narrow size distribution, high saturation magnetization and good morphology control.

3 Functionalization of IONPs One of the most important topics in the design of IONPs for in  vivo applications is functionalization, which provides NPs with high stability in physiological media, stealth and vector targeting properties. In this section, we summarize the most relevant methods to functionalize IONPs for clinical purposes. 3.1 Organic Supra‑structures In recent decades, a class of highly branched and monodispersed macromolecules with well-defined three-dimensional (3D) architectures, such as nanomicelles,

Reprinted from the journal

63

13



Topics in Current Chemistry (2020) 378:40

dendrimers, liposomes and nanogels, have been developed to create hybrid nanoscale materials for imaging and therapeutic applications. 3.1.1 Nanomicelles Nanomicelles are formed by the self-assembly of surfactant molecules or copolymers that adopt a core–shell like structure, thus entrapping in their inner core hydrophobic materials, such as drugs, dyes or inorganic nanoparticles (Fig. 4). The small size is another advantage of the micelles, which can be synthesized between 5 and 100 nm. This provides nanomicelles with long blood circulation times, which favor their active or passive accumulation in the target sites. Consequently, nanomicelles are generating great interest in the development of promising payload nanocarriers for theranostics [217–219]. Particularly interesting are the results obtained with hybrid nanosystems using polymer micelles loaded with IONPs. For instance, Jianping Bin and coworkers described the synthesis of a tumor-targeted MRI vehicle through the encapsulation of IONPs in self-aggregating polymeric folate-conjugated N-palmitoyl chitosan micelles [220]. In vitro and in vivo studies demonstrated the efficacy of folate-conjugated superparamagnetic iron oxide nanoparticle (SPION)micelles in targeting and visualization by MRI of folate receptor overexpressed tumor cells. Torchilin et  al. [221] created a diagnostic and therapeutic agent for in  vivo use based on poly (ethylene glycol)-phosphatidylethanolamine (PEG-PE) micelles loaded with Paclitaxel (PTX), a poorly water soluble anticancer drug, and IONPs. The combination of both multi-modal cargos inside the micelles showed no property changes, either in the relaxivity of the IONPs or in the apoptotic antitumour activity of PTX. 3.1.2 Dendrimers Dendrimers are a class of well-defined nanostructured macromolecules consisting of three critical architectural domains: the multivalent surface, the interior shells surrounding the core, and the core. These domains can be tailored for a specific

Fig. 4  Synthesis of chitosan derivative polymeric micelles encapsulating superparamagnetic iron oxide nanoparticles (SPIONs) [308]

13

64

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

purpose, such as a dendritic sensor [222, 223] or a payload carrier, the encapsulation of molecules in their interior shell being the most used application of dendrimers [224, 225]. One of the most common dendrimers is based on the chemical structure poly (amidoamine) (PAMAM), which has a large number of reactive amine groups on the periphery, making them an excellent platform to construct nanomaterials for biomedical applications [226–228]. Luong et al. [229] designed a promising theranostic agent based on the combination of IONPs and a hydrophobic anticancer drug loaded in a PAMAM dendrimer decorated with folic acid (FA). The design of this hybrid theranostic agent starts with functionalization of the SPIONs with activated carboxyl groups that bind folic acid-PAMAM dendrimers. The engineered SPIONs@FA-PAMAM showed great potential as MRI diagnostic agents, with increased internalization in cancer cells and better image contrast. Moreover, the encapsulation of hydrophobic anticancer drugs, such as 3,4-difluorobenzylidenecurcumin (CDF), in the dendrimers of the SPIONs@FA-PAMAM, enhances their anticancer activity by delivering a higher dose of CDF with high specificity to target cancer cells expressing folate receptors. Dendrimers could also be used in gene therapy as gene delivery platforms. Xiao et al. [220] synthesized a nanohybrid dendrimer based on the combination of PAMAM dendrimers and IONPs through electrostatic interactions. First, the IONPs were functionalized with negatively charged polystyrene sulfonate (PSS), and then positively charged PAMAM dendrimers decorated with plasmid DNA were deposited onto the PSS-functionalized NPs, resulting in a nanohybrid material, PAMAM dendrimer/pDNA-coated MNPs. The results demonstrated that the efficiency of this hybrid system to transfect NIH 3T3 cells is strongly dependent on the dendrimer generation, the amine/phosphate groups ratio and the plasmid DNA concentration. 3.1.3 Liposomes Liposomes comprise a lipid bilayer surrounding an aqueous core. They can be made from different lipid formulation and present different sizes depending on the method of preparation. Similarly to the organic macro-structures mentioned above, liposomes are able to encapsulate payloads in their hydrophobic or hydrophilic inner, which makes them excellent nanocarriers for therapeutic and imaging applications. Liposomes based on phospholipids are the most common vesicles for in vivo applications due to their great advantages, such as biocompatibility, biodegradability and reduced toxicity [230–232]. The incorporation of IONPs into liposomes is gaining increased attention of researchers as a way to synthetize more effective magnetic nanocarriers for in  vivo applications. Di Corato et  al. [233] designed a liposome formulation based on phosphatidylcholine lipids that entraps magnetic NPs and a photosensitizer in its interior. In a single synthesis method, higher concentrations of hydrophilic IONPs were encapsulated in the core, and a hydrophobic photosynthesizer, Temoporfin (marketed as Foscan), was incorporated into the lipid bilayer. The resulting magnetic liposome presented double functionality, magnetic hyperthermia and photodynamic therapy, which led to complete death of cancer cells in vitro and total ablation of solid-tumor in vivo. Reprinted from the journal

65

13

Topics in Current Chemistry (2020) 378:40

Zheng et al. [234] synthetized a tumor-specific peptide-decorated liposome containing payloads of IONPs and an anti-cancer drug in their inner core and lipid bilayer, respectively. Like the protocol described above, the combination in a single pot reaction of egg phosphatidylcholine, cholesterol, paclitaxel (PTX), different 1,2-distearoyl-snglycero-3-phosphoethanolamine (DSPE) phospholipids, such as DSPE-PEG and cell penentrating peptide-modified DSPE-PEG, and hydrophilic SPION, generated a theranostic liposome. The results confirmed the effectiveness for tumor targeting and antitumor activity through MRI in vivo experiments. 3.1.4 Nanogels Nanogels (NGs) are nanosized water-soluble particles formed by crosslinked polymer networks with loading capacity of therapeutics. Stimuli-responsive NGs are a class of smart particles that respond to external physical changes, such as pH, temperature or redox agents [235, 236]. This behavior allows the controlled-release of payloads from NGs, minimizing possible side effects and avoiding the use of high doses. NGs can also be loaded with diagnostic agents, such as magnetic NPs, enabling their visualization and follow-up by MRI. These characteristics, together with the ease of uptake by cancer cells and tumor tissues due to their softness and fluidity, make NG-based nanosystems a high potential theranostic material [237, 238]. Qian et al. [239] prepared a hybrid NG system based on a thermo-responsive copolymer [N-isopropylacrylamide, methacrylic acid and poly (ethylene glycol) methacrylate] that stabilizes hydrophobic IONPs and 10-hydroxy camptothecin (HCPT) in its inner compartment. The obtained IONP/HCPT-NG generated an increase in reactive oxygen species (ROS), allowed the enrichment of NG at the tumor site by applying an external magnetic field, and offered the possibility of being used as nanocarrier for photothermal therapy due to its absorption in the near infrared (NIR) range. In vivo results demonstrated that the combination of PTT and chemotherapy with external magnetic fields on IONP/HCPT-NGs, reduced the growth of primary tumors and prevented metastasis [239]. Alginate (AG) is a natural polysaccharide that has been gaining attraction in recent years for the synthesis of polymeric nanomaterials with biomedical applications thanks to its biocompatibility, biodegradability and ease of gelation [240]. For instance, Hao et  al. [241] designed alginate NGs loaded with IONPs and bone mesenchymal stem cells (BMSCs) for enhanced tumour MR imaging (Fig. 5). The potential advantage of using BMSCs as tumor delivery vehicles is that they are not tumorigenic and minimally immunogenic. In this way, polyethilenimine (PEI)-functionalized IONPs were crosslinked to AG NGs previously synthesized by a double emulsion method. The resulting AG/PEI-NP NGs were taken up by BMSCs without affecting cell characteristics. BSMC-AG/PEI-NP NGs were then used successfully for the in vivo diagnosis of different tumor models [241].

13

66

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



Fig. 5  Schematic illustration of alginate/polyethilenimine-iron (III) oxide (AG/PEI-Fe3O4) and stem-cellmediated delivery of nanogels (NGs) for enhanced breast or glioma tumor molecular resonance (MR) imaging. Reprinted with permission from [241]. Copyright (2019) Royal Society of Chemistry

3.2 Inorganic Coverage Mesoporous silica is the most important inorganic coating material for IONPs due to the ease of functionalization, high stability, and vast surface area and pore volume to host large number molecules. These characteristics make hybrid mesoporous silicaIONPs excellent nanocarriers for controlled drug release therapies [242, 243]. Based on this, Vallet-Regí et al. [244] designed a responsive silica matrix nanocarrier for tumor therapy based on magnetic NPs that combine the heat release mediated by magnetic hyperthermia and doxorubicin release through a thermoresponsive polymer. The as-prepared OA-capped IONPs are transferred into aqueous solution with CTAB, which helps the growth of the silica matrix by addition of tetraethyl orthosilicate (TEOS) as a silica precursor. Then, the silica-matrix-coated IONPs are functionalized with a methacrylate molecule as a polymer precursor to perform, using N-isopropylacrylamide (NIPAM), N-(hydroxymethyl)acrylamide (NHMA), and N,N′-methylenebis(acrylamide) (MBA) monomers, the synthesis of a thermoresponsive polymer surrounding the mesoporous silica-coated IONP (Fig. 6). Direct injection into the tumor site of Doxo-loaded mesoporous silica NPs, together with the application of amplified magnetic fields, provoked a synergistic effect between magnetic hyperthermia and chemotherapy that led to significant tumor growth inhibition and low toxicity [244]. Hurley et al. [245] demonstrated that the inclusion of functionalized mesoporous silica coating in IONPs cores results in stable NPs with high heat capacity and high MRI contrast. The anionic surfactants capped IONPs (a commercially available IONP called EMG-308) required pre-functionalization with polyvinylpyrrolidone (PVP) prior to silica condensation with the TEOS precursor. Finally, the functionalization with PEG and trimethyl silane derivates yields colloidal stable NPs with the same magnetic character that un-functionalized IONPs and minimal toxicity toward human skin fibroblasts. Furthermore, a direct injection into LNCaP prostate cancer tumours implanted in nude mice showed that these hybrid mesoporous silica-IONPs can improve the heating and imaging contrast of IONPs [245]. Reprinted from the journal

67

13



Topics in Current Chemistry (2020) 378:40

Fig. 6  Synthesis of mesoporous silica-coated (ms)-IONPs. Polyvinylpyrrolidone (PVP)-10 was added to IONPs prior to cetyltrimethylammonium bromide (CTAB) addition and silica condensation to allow for CTAB colocalization with IONPs and to maintain a spacer layer between the silica shell and IONP core. Reprinted with permission from [245]. Copyright (2016) American Chemical Society

3.3 Ligand Exchange Ligand exchange is a very complicated coating strategy that involves multiple interactions potencials/forces. It requires the use of reactive binding molecules that enable the replacement of capping agents attached to the nanoparticle surfaces. This binding between the iron atoms of the IONP and the anchor group of the ligand molecules is mediated by electrostatic interactions. Therefore, the nature of the anchor group is determinant in the search for highly stable ligand molecules at the IONP surfaces. In addition to anchor groups, the hydrophilic balance of the ligand is also important to render water-soluble NPs [236, 246, 247]. In our group, we have developed different ligand formulations to functionalize IONPs to obtain soluble and stable NPs in physiological media for in vivo MRI applications. These ligands are based on a gallol group as a strong binder and PEG chains as hydrophilic tunable spacers, which also minimize plasma protein adsorption. In this manner, we have demonstrated that selection of the right molecular weight of PEG chain and the outermost charged group of the ligand plays a fundamental role in the fate and bioavailability of intravenously injected IONPs. Thus, a ligand with a PEG chain between 1500 and 3000  Da and neutral outermost groups showed the best stealth properties, resulting in longer blood circulation times and higher bioavailability without increased toxicity [248–250].

4 Applications of IONPs in MRI Among the main clinical diagnostic techniques, MRI stands out for its unique combination of qualities, such as its non-invasive character, the absence of ionizing radiation, excellent image quality, and its ability to provide both anatomical and functional information [251]. The MRI signal comes mainly from the protons of the water molecules, while the image contrast is generated from differences in the intensity of this signal among different tissues, which depends on the concentration, relaxation times ­(T1 and ­T2) and mobility of the water molecules within each tissue [252, 253]. Additionally, image contrast can be further enhanced using CAs. Although there are several mechanisms that can produce MRI contrast, such as

13

68

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



chemical exchange saturation transfer (CEST) or hyperpolarization, most MRI CAs produce contrast by altering the relaxation times of the surrounding water protons [254, 255]. The capacity of a CA to decrease the relaxation times ­(T1 or ­T2) is given by a parameter known as relaxivity ­(r1 or ­r2), which is expressed in ­mM−1·s−1. In MRI, among the most commonly used CA are chelates of paramagnetic gadolinium(III) ions ­(Gd3+). However, conventional Gd-chelates have some important limitations, such as the lack of diagnostic specificity and the toxicity associated with their use as a result of the unexpected release of free Gd ions [256, 257]. Magnetic NPs have emerged as a promising alternative to overcome these limitations [258]. 4.1 IONPs in Tumor Diagnosis 4.1.1 Untargeted IONPs The evaluation of IONPs as CAs in cancer research is performed mainly in rodent models, called ‘indirect xenografts’ [259]. Cancer cells can be implanted either into a tissue unrelated to the original tumor site (heterotopic model) or into the corresponding anatomical position (orthotopic model) [260] (Fig. 7). The route of administration of magnetic NPs is also relevant as it influences the biodistribution and pharmacokinetics of the CA. Several administration routes have been used in preclinical studies, mainly intratumoral, intraperitoneal or intravenous injection; for obvious reasons, the latter is the most interesting for clinical applications. After intravenous administration, IONPs have been described to accumulate in tumors due to the EPR (Enhanced Permeability and Retention) effect. This passive transport is determined by the high vascularization of tumors, and therefore increased blood flow, together with increased vascular permeability and poor lymphatic drainage [261]. Efremova et al. [262] developed IONPs for diagnosis of breast cancer in a heterotopic model. They observed that IONPs accumulated passively inside the tumor 24 h after intravenous injection using T ­ 2-weighted MR images. Similar studies have been conducted using orthotopic models of breast cancer [263, 264], pancreatic cancer [265] and glioblastoma multiforme (GBM) [266]. All these studies conclude that IONPs accumulated in the tumor due to the EPR effect; however, most of them lack quantitative analyses, which are necessary to determine the amount of IONPs that actually reach the tumor. Intratumoral administration could be an alternative for tumor therapy when the CA is not able to reach the tumor by a venous route. However, this approach has serious limitations for diagnostic applications since, in most cases, it would not add any useful information to that already provided by the MR images without CA. Furthermore, intratumoral administration makes no sense when it comes to very early diagnosis, detection of metastasis or in the case of inaccessible tumors. Nevertheless, several preclinical studies have been conducted using intratumoral injection of IONPs [267–269]. The authors used qualitative MRI to evaluate the distribution of IONPs throughout the tumor, which showed that IONPs spread slowly and inefficiently. Therefore, in these studies the information provided by MRI after the Reprinted from the journal

69

13



Topics in Current Chemistry (2020) 378:40

Fig. 7  C6 brain tumor model implanted orthotopically (upper panels) and heterotopically (lower panels). Left) ­T2-weighted MR images before the injection of IONPs; right) ­T2-weighted MR images 1 h after the injection of IONPs

intratumoral injection of IONPs serves as proof of concept, but, as we have just mentioned, it is of no practical value for potential clinical applications. In conclusion, up to now, untargeted IONPs have not proven to be a good alternative to conventional MRI CAs for cancer diagnosis. 4.1.2 Targeted IONPs To improve the accumulation of IONPs in tumors, a promising strategy is conjugation with targeting segments [5]. In principle, this functionalization would allow not only the visualization of IONPs by MRI, but would also offer the possibility of visualizing cellular and subcellular functions and processes in living organisms without perturbing them, giving rise to so-called molecular MRI (mMRI) [270], which was first described by Richard Klausner [271, 272] (Fig. 8).

13

70

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



Fig. 8  Scheme of the non-targeted (top) and targeted IONPs (bottom)

It is worth mentioning that the targeted strategy for cancer diagnosis was described before the untargeted strategy. Reimer et al. [273] described in 1990 the diagnosis of liver cancer after the intravenous administration of IONPs-arabinogalactan conjugates, in both heterotopic and orthotopic models. IONPs have been functionalized with epithelial growth factor (EGFR) antibodies for the diagnosis of breast cancer [274], pancreatic/stomach cancer [275] and brain cancer [276]. Because there is an established relationship between mutations involving overexpression or overactivity of EGFR and various types of cancer, this receptor is currently one of the most important targets in cancer research [277–279]. Similarly, PSCA (prostate stem cell antigen) antibody was bound to IONPs for diagnosis of prostate cancer [280]. Integrins receptor, especially α5β3, has been found to be differentially overexpressed in tumors, playing a vital role in tumor angiogenesis [281–283]. Integrins are recognized mainly by short peptide sequences, such as Arg–Gly–Asp (RGD). Therefore, some NPs functionalized with RGD have been proposed for the diagnosis of brain cancer [284], colon cancer [285] or fibrosarcoma [286], among others. Among other functionalization molecules for targeted diagnosis, it is worth highlighting the use of aptamers for kidney [287] and liver cancer [288], peptides for prostate and liver cancer [289, 290], and flavin adenine dinucleotide (FAD) for prostate cancer [291].

Reprinted from the journal

71

13

Topics in Current Chemistry (2020) 378:40

Fig. 9  a In vivo MR images of a NCr nude mouse at different time points after intravenous injection of IONPs. b Quantification of liver contrast collected at different time points after accumulation of IONPs in NCr nude mice. c In  vivo MR images of liver tumor orthotopic xenographs at different time points after intravenous injection of IONPs. d Quantification of contrast-to-noise ratio (CNR) of tumor-to-liver contrast at different time points. e, f Histopathological analysis of mouse liver 1 h after the intravenous injection of IONPs. Reprinted with permission from [292]. Copyright (2018) American Chemical Society

Recently, Chee et al. published an interesting study in which they described the design of a library of short peptides and ligands to functionalize IONPs. From this library, they selected the ligand that provided IONPs with the best characteristics for in  vivo use, namely, long term stability, non-specific binding to live cells and absence of cytotoxicity at high concentrations. IONPs functionalized with this ligand showed a significant increase in contrast between the liver tumor and the healthy liver tissue, as compared with commercial MRI CAs [292] (Fig. 9). Finally, it is worth mentioning that, in clinical diagnosis, positive contrast is generally preferred over negative contrast because it avoids the potential confusion of signal decay caused by negative CAs with signal voids caused by magnetic field inhomogeneities induced by air, metal prosthesis, etc. Thus, a very recent study described the use of Cu as a dopant agent that enhances the positive contrast of IONPs functionalized with RGD for targeted diagnosis of breast cancer [184]. Even though many IONPs show dual contrast potential, that is, ­r2/r1 ratio  between 3 and 10, their use in vivo as positive CAs is limited by the acquisition conditions of conventional ­T1-weighted MRI sequences, which are usually based on the spin-echo acquisition scheme and therefore require relatively long echo times. However, the introduction of new MRI acquisition sequences, such as ultra-short echo time (UTE) sequences, is making it possible to detect IONPs as positive contrast [293].

13

72

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



4.2 IONPs as CA in Other Pathologies Although most research in IONPs designed to serve as MRI CAs is focused on cancer diagnosis, there are many other pathologies that can benefit from advances in this field of research, as discussed henceforth. Recent investigations have demonstrated that an acetylcholine-sensitive mMRI nanosensor can be used for measuring the endogenous release of acetylcholine in the rat brain after its intracerebral administration [294]. Similarly, after intracerebral administration of alginate-coated IONPs, changes in C ­ a2+ levels have been monitored following a quinolinic acid-induced striatal lesion [295]. Other magnetic nanostructures have been used for the detection of brain inflammation [296]. These particular nanostructures are based on IONPs coupled covalently through peptide linkers that have been designed to be cleaved by the intracellular macrophage cathepsin, which results in microparticles of iron oxide (MPIO) and allows the fate of magnetic NPs to be tracked. This is because the MPIO, once sequestered by macrophages in the liver, decrease their relaxivity, while particles that associate with their target tissue in the brain remain unaltered and functional. Thrombosis is a major clinical problem whose incidence has not decreased over the last 20  years and is involved in several pathological disorders such as myocardial infarction, ischemic stroke or pulmonary embolism, among others [297]. Early detection is essential for effective treatment, but it remains challenging in practice. P-selectin is an adhesion molecule, overexpressed at the surface of endothelial cells and platelets upon activation, which plays a fundamental role in thrombus formation [298]. Based on this fact, Suzuki et  al. [299] innovated a fucoidan (a natural sulfated polysaccharide with high affinity for activated platelets through P-selectin)-coated USPIONs to visualize by MRI arterial thrombi in the early stage of the disease. Other investigations used PLGA-coated IONs, functionalized with EWVDV peptide, which has a high affinity and specificity for P-selectin, to target thrombi for both diagnosis and treatment through the induction of thrombolysis [300]. 4.3 Other Applications IONPs have also been used in combination with MRI for many other in vivo applications, such as imaging of activated microglia during brain inflammation [301], tracking of stem cells [302–304], image-guided treatment of anemia using bacteria loaded with IONPs [305], or to carry out vascular imaging [306, 307], among others.

Reprinted from the journal

73

13

Topics in Current Chemistry (2020) 378:40

5 Conclusions Recent advances in nanotechnology applied to biomedical research have made possible the development of a new generation of magnetic nanomaterials with great potential as MRI CAs. IONPs stand out due to their excellent combination of properties for in vivo applications, that is, their superparamagnetism along with their high biocompatibility. Also, advanced functionalization strategies have allowed these IONPs to be specifically targeted to different tissues or cells to perform molecular imaging. However, in spite of all these advances, and the large number of studies carried out in this field, very little clinical translation has been achieved so far. The main reasons behind this relative failure are very likely related to reproducibility and scalability issues during the synthesis process, which must be further improved. Also, in vivo studies must be thoroughly designed to include comprehensive toxicity assays and preclinical imaging studies using appropriate animal models. Acknowledgements  Ashish Avasthi thanks the Marie Curie COFUND program for her PhD scholarship (NanoMedPhD, Grant agreement 713721). Financial support was provided by the Spanish Ministry of Science, Innovation and Universities (CTQ2017-86655-R) to María Luisa García-Martín and Manuel Pernia Leal. Manuel Pernia Leal also thanks the “V Plan Propio” of the University of Seville for his Postdoctoral Fellowship.

References 1. García-Martín ML, López-Larrubia P (2018) Preclinical MRI: methods and protocols, vol 1718. Methods in molecular biology. Springer, New York 2. Xie W, Guo Z, Gao F, Gao Q, Wang D, Liaw B-s, Cai Q, Sun X, Wang X, Zhao L (2018) Shape-, size- and structure-controlled synthesis and biocompatibility of iron oxide nanoparticles for magnetic theranostics. Theranostics 8(12):3284–3307. https​://doi.org/10.7150/thno.25220​ 3. Leong HS, Butler KS, Brinker CJ, Azzawi M, Conlan S, Dufés C, Owen A, Rannard S, Scott C, Chen C, Dobrovolskaia MA, Kozlov SV, Prina-Mello A, Schmid R, Wick P, Caputo F, Boisseau P, Crist RM, McNeil SE, Fadeel B, Tran L, Hansen SF, Hartmann NB, Clausen LPW, Skjolding LM, Baun A, Ågerstrand M, Gu Z, Lamprou DA, Hoskins C, Huang L, Song W, Cao H, Liu X, Jandt KD, Jiang W, Kim BYS, Wheeler KE, Chetwynd AJ, Lynch I, Moghimi SM, Nel A, Xia T, Weiss PS, Sarmento B, das Neves J, Santos HA, Santos L, Mitragotri S, Little S, Peer D, Amiji MM, Alonso MJ, Petri-Fink A, Balog S, Lee A, Drasler B, Rothen-Rutishauser B, Wilhelm S, Acar H, Harrison RG, Mao C, Mukherjee P, Ramesh R, McNally LR, Busatto S, Wolfram J, Bergese P, Ferrari M, Fang RH, Zhang L, Zheng J, Peng C, Du B, Yu M, Charron DM, Zheng G, Pastore C (2019) On the issue of transparency and reproducibility in nanomedicine. Nat Nanotechnol 14(7):629–635. https​://doi.org/10.1038/s4156​5-019-0496-9 4. Wu W, He Q, Jiang C (2008) Magnetic iron oxide nanoparticles: synthesis and surface functionalization strategies. Nanosc Res Lett 3(11):397–415. https​://doi.org/10.1007/s1167​1-008-9174-9 5. Zhi D, Yang T, Yang J, Fu S, Zhang S (2020) Targeting strategies for superparamagnetic iron oxide nanoparticles in cancer therapy. Acta Biomater 102:13–34. https​://doi.org/10.1016/j.actbi​ o.2019.11.027 6. Ventola CL (2017) Progress in nanomedicine: approved and investigational nanodrugs. Pharm Therap 42(12):742–755 7. Massart R (1981) Preparation of aqueous magnetic liquids in alkaline and acidic media. IEEE Trans Magn 17(2):1247–1248. https​://doi.org/10.1109/TMAG.1981.10611​88 8. Le Fort J (1852) C R Acad Sci Paris 34:480

13

74

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40 9. Elmore WC (1938) The magnetization of ferromagnetic colloids. Phys Rev 54(12):1092–1095. https​://doi.org/10.1103/PhysR​ev.54.1092 10. LaMer VK, Dinegar RH (1950) Theory, production and mechanism of formation of monodispersed hydrosols. J Am Chem Soc 72(11):4847–4854. https​://doi.org/10.1021/ja011​67a00​1 11. Banfield JF, Welch SA, Zhang H, Ebert TT, Penn RL (2000) Aggregation-based crystal growth and microstructure development in natural iron oxyhydroxide biomineralization products. Science 289(5480):751–754. https​://doi.org/10.1126/scien​ce.289.5480.751 12. Boistelle R, Astier JP (1988) Crystallization mechanisms in solution. J Cryst Growth 90(1):14–30. https​://doi.org/10.1016/0022-0248(88)90294​-1 13. Penn RL, Banfield JF (1998) Imperfect oriented attachment: dislocation generation in defect-free nanocrystals. Science 281(5379):969–971. https​://doi.org/10.1126/scien​ce.281.5379.969 14. Penn RL, Banfield JF (1999) Morphology development and crystal growth in nanocrystalline aggregates under hydrothermal conditions: insights from titania. Geochim Cosmochim Acta 63(10):1549–1557. https​://doi.org/10.1016/S0016​-7037(99)00037​-X 15. LaGrow AP, Besenhard MO, Hodzic A, Sergides A, Bogart LK, Gavriilidis A, Thanh NTK (2019) Unravelling the growth mechanism of the co-precipitation of iron oxide nanoparticles with the aid of synchrotron X-ray diffraction in solution. Nanoscale 11(14):6620–6628. https​://doi.org/10.1039/ c9nr0​0531e​ 16. Bee A, Massart R, Neveu S (1995) Synthesis of very fine maghemite particles. J Magn Magn Mater 149(1):6–9. https​://doi.org/10.1016/0304-8853(95)00317​-7 17. Babes L, Denizot B, Tanguy G, Le Jeune JJ, Jallet P (1999) Synthesis of iron oxide nanoparticles used as MRI contrast agents: a parametric study. J Colloid Interface Sci 212(2):474–482. https​:// doi.org/10.1006/jcis.1998.6053 18. Qiu X-P (2000) Synthesis and characterization of magnetic nano particles. Chin J Chem 18(6):834–837. https​://doi.org/10.1002/cjoc.20000​18060​7 19. da Costa GM, De Grave E, de Bakker PMA, Vandenberghe RE (1994) Synthesis and characterization of some iron oxides by sol–gel method. J Solid State Chem 113(2):405–412. https​://doi. org/10.1006/jssc.1994.1388 20. Sjogren CE, Briley-Saebo K, Hanson M, Johansson C (1994) Magnetic characterization of iron oxides for magnetic resonance imaging. Magn Reson Med 31(3):268–272. https​://doi.org/10.1002/ mrm.19103​10305​ 21. Gupta AK, Curtis AS (2004) Lactoferrin and ceruloplasmin derivatized superparamagnetic iron oxide nanoparticles for targeting cell surface receptors. Biomaterials 25(15):3029–3040. https​:// doi.org/10.1016/j.bioma​teria​ls.2003.09.095 22. Kim DK, Zhang Y, Voit W, Rao KV, Muhammed M (2001) Synthesis and characterization of surfactant-coated superparamagnetic monodispersed iron oxide nanoparticles. J Magn Magn Mater 225(1):30–36. https​://doi.org/10.1016/S0304​-8853(00)01224​-5 23. Vayssières L, Chanéac C, Tronc E, Jolivet JP (1998) Size tailoring of magnetite particles formed by aqueous precipitation: an example of thermodynamic stability of nanometric oxide particles. J Colloid Interface Sci 205(2):205–212. https​://doi.org/10.1006/jcis.1998.5614 24. Jiang W, Yang HC, Yang SY, Horng HE, Hung JC, Chen YC, Hong C-Y (2004) Preparation and properties of superparamagnetic nanoparticles with narrow size distribution and biocompatible. J Magn Magn Mater 283(2):210–214. https​://doi.org/10.1016/j.jmmm.2004.05.022 25. Tominaga M, Matsumoto M, Soejima K, Taniguchi I (2006) Size control for two-dimensional iron oxide nanodots derived from biological molecules. J Colloid Interface Sci 299(2):761–765. https​:// doi.org/10.1016/j.jcis.2006.02.022 26. Itoh H, Sugimoto T (2003) Systematic control of size, shape, structure, and magnetic properties of uniform magnetite and maghemite particles. J Colloid Interface Sci 265(2):283–295. https​://doi. org/10.1016/S0021​-9797(03)00511​-3 27. Pardoe H, Chua-anusorn W, St. PierreDobson TGJ (2001) Structural and magnetic properties of nanoscale iron oxide particles synthesized in the presence of dextran or polyvinyl alcohol. J Magn Magn Mater 225(1):41–46. https​://doi.org/10.1016/S0304​-8853(00)01226​-9 28. Thapa D, Palkar VR, Kurup MB, Malik SK (2004) Properties of magnetite nanoparticles synthesized through a novel chemical route. Mater Lett 58(21):2692–2694. https​://doi.org/10.1016/j. matle​t.2004.03.045 29. Khalafalla S, Reimers G (1980) Preparation of dilution-stable aqueous magnetic fluids. IEEE Trans Magn 16(2):178–183. https​://doi.org/10.1109/TMAG.1980.10605​78

Reprinted from the journal

75

13



Topics in Current Chemistry (2020) 378:40

30. Jolivet J-P, Froidefond C, Pottier A, Chanéac C, Cassaignon S, Tronc E, Euzen P (2004) Size tailoring of oxide nanoparticles by precipitation in aqueous medium. A semi-quantitative modelling. J Mater Chem 14(21):3281–3288. https​://doi.org/10.1039/B4070​86K 31. Bui TQ, Ton SN-C, Duong AT, Tran HT (2018) Size-dependent magnetic responsiveness of magnetite nanoparticles synthesised by co-precipitation and solvothermal methods. J Sci Adv Mater Devices 3(1):107–112. https​://doi.org/10.1016/j.jsamd​.2017.11.002 32. Thomas JR (1966) Preparation and magnetic properties of colloidal cobalt particles. J Appl Phys 37(7):2914–2915. https​://doi.org/10.1063/1.17821​54 33. Smith TW, Wychick D (1980) Colloidal iron dispersions prepared via the polymer-catalyzed decomposition of iron pentacarbonyl. J Phys Chem 84:1621–1629. https​://doi.org/10.1021/j1004​ 49a03​7 34. Hess PH, Parker PH Jr (1966) Polymers for stabilization of colloidal cobalt particles. J Appl Polym Sci 10(12):1915–1927. https​://doi.org/10.1002/app.1966.07010​1209 35. Peter K, Vollhardt C, Bercaw JE, Bergman RG (1975) Photochemistry of η5 cylclopentadienylcobalt) tricarbonyl, tris(η5-cyclopentadienylcobalt monocarbonyl) and tetra(η5cyclopentadienylcobalt) dicarbonyl*. J Organomet Chem 97(2):283–297. https​://doi.org/10.1016/ S0022​-328X(00)89475​-9 36. Van Wonterghem J, Mørup S, Charles SW, Wells S (1988) An investigation of the chemical reactions leading to the formation of ultrafine amorphous fe100−xcx alloy particles. J Colloid Interface Sci 121(2):558–563. https​://doi.org/10.1016/0021-9797(88)90457​-2 37. Hyeon T, Lee SS, Park J, Chung Y, Na HB (2001) Synthesis of highly crystalline and monodisperse maghemite nanocrystallites without a size-selection process. J Am Chem Soc 123(51):12798– 12801. https​://doi.org/10.1021/ja016​812s 38. Jana NR, Chen Y, Peng X (2004) Size- and shape-controlled magnetic (Cr, Mn, Fe Co, Ni) oxide nanocrystals via a simple and general approach. Chem Mater 16(20):3931–3935. https​://doi. org/10.1021/cm049​221k 39. Lassenberger A, Grünewald TA, van Oostrum PDJ, Rennhofer H, Amenitsch H, Zirbs R, Lichtenegger HC, Reimhult E (2017) Monodisperse iron oxide nanoparticles by thermal decomposition: elucidating particle formation by second-resolved in situ small-angle X-ray scattering. Chem Mater 29(10):4511–4522. https​://doi.org/10.1021/acs.chemm​ater.7b012​07 40. Kwon SG, Piao Y, Park J, Angappane S, Jo Y, Hwang NM, Park JG, Hyeon T (2007) Kinetics of monodisperse iron oxide nanocrystal formation by "heating-up" process. J Am Chem Soc 129(41):12571–12584. https​://doi.org/10.1021/ja074​633q 41. Kim BH, Shin K, Kwon SG, Jang Y, Lee H-S, Lee H, Jun SW, Lee J, Han SY, Yim Y-H, Kim D-H, Hyeon T (2013) Sizing by weighing: characterizing sizes of ultrasmall-sized iron oxide nanocrystals using MALDI-TOF mass spectrometry. J Am Chem Soc 135(7):2407–2410. https​:// doi.org/10.1021/ja310​030c 42. Kura H, Takahashi M, Ogawa T (2010) Synthesis of monodisperse iron nanoparticles with a high saturation magnetization using an Fe(CO)x−oleylamine reacted precursor. J Phys Chem C 114(13):5835–5838. https​://doi.org/10.1021/jp911​161g 43. Sun S, Zeng H, Robinson DB, Raoux S, Rice PM, Wang SX, Li G (2004) Monodisperse M ­ Fe2O4 (M = Fe Co, Mn) nanoparticles. J Am Chem Soc 126(1):273–279. https​://doi.org/10.1021/ja038​ 0852 44. Rockenberger J, Scher EC, Alivisatos AP (1999) A new nonhydrolytic single-precursor approach to surfactant-capped nanocrystals of transition metal oxides. J Am Chem Soc 121(49):11595–11596. https​://doi.org/10.1021/ja993​280v 45. Park J, An K, Hwang Y, Park JG, Noh HJ, Kim JY, Park JH, Hwang NM, Hyeon T (2004) Ultralarge-scale syntheses of monodisperse nanocrystals. Nat Mater 3(12):891–895. https​://doi. org/10.1038/nmat1​251 46. Park S-J, Kim S, Lee S, Khim ZG, Char K, Hyeon T (2000) Synthesis and magnetic studies of uniform iron nanorods and nanospheres. J Am Chem Soc 122(35):8581–8582. https​://doi.org/10.1021/ ja001​628c 47. Zhang H, Li L, Liu XL, Jiao J, Ng C-T, Yi JB, Luo YE, Bay B-H, Zhao LY, Peng ML, Gu N, Fan HM (2017) Ultrasmall ferrite nanoparticles synthesized via dynamic simultaneous thermal decomposition for high-performance and multifunctional T1 magnetic resonance imaging contrast agent. ACS Nano 11(4):3614–3631. https​://doi.org/10.1021/acsna​no.6b076​84

13

76

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40 48. Roca AG, Morales MP, Serna CJ (2006) Synthesis of monodispersed magnetite particles from different organometallic precursors. IEEE Trans Magn 42(10):3025–3029. https​://doi.org/10.1109/ TMAG.2006.88011​1 49. Guardia P, Riedinger A, Nitti S, Pugliese G, Marras S, Genovese A, Materia ME, Lefevre C, Manna L, Pellegrino T (2014) One pot synthesis of monodisperse water soluble iron oxide nanocrystals with high values of the specific absorption rate. J Mater Chem B 2(28):4426–4434. https​://doi.org/10.1039/C4TB0​0061G​ 50. Kim BH, Lee N, Kim H, An K, Park YI, Choi Y, Shin K, Lee Y, Kwon SG, Na HB, Park J-G, Ahn T-Y, Kim Y-W, Moon WK, Choi SH, Hyeon T (2011) Large-scale synthesis of uniform and extremely small-sized iron oxide nanoparticles for high-resolution T1 magnetic resonance imaging contrast agents. J Am Chem Soc 133(32):12624–12631. https​://doi.org/10.1021/ja203​340u 51. Kovalenko MV, Bodnarchuk MI, Lechner RT, Hesser G, Schäffler F, Heiss W (2007) Fatty acid salts as stabilizers in size- and shape-controlled nanocrystal synthesis: the case of inverse spinel iron oxide. J Am Chem Soc 129(20):6352–6353. https​://doi.org/10.1021/ja069​2478 52. Mourdikoudis S, Liz-Marzán LM (2013) Oleylamine in nanoparticle synthesis. Chem Mater 25(9):1465–1476. https​://doi.org/10.1021/cm400​0476 53. Hou Y, Xu Z, Sun S (2007) Controlled synthesis and chemical conversions of FeO nanoparticles. Angew Chem Int Ed 46(33):6329–6332. https​://doi.org/10.1002/anie.20070​1694 54. Nemati Z, Alonso J, Rodrigo I, Das R, Garaio E, García JÁ, Orue I, Phan M-H, Srikanth H (2018) Improving the heating efficiency of iron oxide nanoparticles by tuning their shape and size. J Phys Chem C 122(4):2367–2381. https​://doi.org/10.1021/acs.jpcc.7b105​28 55. Park J, Lee E, Hwang N-M, Kang M, Kim SC, Hwang Y, Park J-G, Noh H-J, Kim J-Y, Park J-H, Hyeon T (2005) One-nanometer-scale size-controlled synthesis of monodisperse magnetic iron oxide nanoparticles. Angew Chem Int Ed 44(19):2872–2877. https​://doi.org/10.1002/anie.20046​ 1665 56. Li Z, Sun Q, Gao M (2005) Preparation of water-soluble magnetite nanocrystals from hydrated ferric salts in 2-pyrrolidone: mechanism leading to F ­ e3O4. Angew Chem Int Ed 44(1):123–126. https​ ://doi.org/10.1002/anie.20046​0715 57. Xu Z, Shen C, Hou Y, Gao H, Sun S (2009) Oleylamine as both reducing agent and stabilizer in a facile synthesis of magnetite nanoparticles. Chem Mater 21(9):1778–1780. https​://doi.org/10.1021/ cm802​978z 58. Yang H, Ogawa T, Hasegawa D, Takahashi M (2008) Synthesis and magnetic properties of monodisperse magnetite nanocubes. J Appl Phys 103(7):07D526. https​://doi.org/10.1063/1.28338​20 59. Mohapatra J, Mitra A, Bahadur D, Aslam M (2013) Surface controlled synthesis of ­MFe2O4 (M = Mn, Fe Co, Ni and Zn) nanoparticles and their magnetic characteristics. CrystEngComm 15(3):524–532. https​://doi.org/10.1039/C2CE2​5957E​ 60. Macdonald JE, Brooks CJ, Veinot JGC (2008) The influence of trace water concentration on iron oxide nanoparticle size. Chem Commun 32:3777–3779. https​://doi.org/10.1039/B8057​15J 61. Unni M, Uhl AM, Savliwala S, Savitzky BH, Dhavalikar R, Garraud N, Arnold DP, Kourkoutis LF, Andrew JS, Rinaldi C (2017) Thermal decomposition synthesis of iron oxide nanoparticles with diminished magnetic dead layer by controlled addition of oxygen. ACS Nano 11(2):2284–2303. https​://doi.org/10.1021/acsna​no.7b006​09 62. Li Z, Chen H, Bao H, Gao M (2004) One-pot reaction to synthesize water-soluble magnetite nanocrystals. Chem Mater 16(8):1391–1393. https​://doi.org/10.1021/cm035​346y 63. Ooi F, DuChene JS, Qiu J, Graham JO, Engelhard MH, Cao G, Gai Z, Wei WD (2015) A facile solvothermal synthesis of octahedral ­Fe3O4 nanoparticles. Small 11(22):2649–2653. https​://doi. org/10.1002/smll.20140​1954 64. Bunge A, Porav AS, Borodi G, Radu T, Pîrnău A, Berghian-Grosan C, Turcu R (2019) Correlation between synthesis parameters and properties of magnetite clusters prepared by solvothermal polyol method. J Mater Sci 54(4):2853–2875. https​://doi.org/10.1007/s1085​3-018-3030-9 65. Mao B, Kang Z, Wang E, Lian S, Gao L, Tian C, Wang C (2006) Synthesis of magnetite octahedrons from iron powders through a mild hydrothermal method. Mater Res Bull 41(12):2226–2231. https​://doi.org/10.1016/j.mater​resbu​ll.2006.04.037 66. Giri J, Guha Thakurta S, Bellare J, Kumar Nigam A, Bahadur D (2005) Preparation and characterization of phospholipid stabilized uniform sized magnetite nanoparticles. J Magn Magn Mater 293(1):62–68. https​://doi.org/10.1016/j.jmmm.2005.01.044

Reprinted from the journal

77

13



Topics in Current Chemistry (2020) 378:40

67. Wang J, Sun J, Sun Q, Chen Q (2003) One-step hydrothermal process to prepare highly crystalline ­Fe3O4 nanoparticles with improved magnetic properties. Mater Res Bull 38(7):1113–1118. https​:// doi.org/10.1016/S0025​-5408(03)00129​-6 68. Li J, Shi X, Shen M (2014) Hydrothermal synthesis and functionalization of iron oxide nanoparticles for mr imaging applications. Part Part Syst Charact 31(12):1223–1237. https​://doi. org/10.1002/ppsc.20140​0087 69. Lassoued A, Lassoued MS, Dkhil B, Ammar S, Gadri A (2018) Synthesis, photoluminescence and magnetic properties of iron oxide (α-Fe2O3) nanoparticles through precipitation or hydrothermal methods. Phys E 101:212–219. https​://doi.org/10.1016/j.physe​.2018.04.009 70. Pinna N, Grancharov S, Beato P, Bonville P, Antonietti M, Niederberger M (2005) Magnetite nanocrystals: nonaqueous synthesis, characterization, and solubility. Chem Mater 17(11):3044– 3049. https​://doi.org/10.1021/cm050​060+ 71. Nassar MY, Ahmed IS, Hendy HS (2018) A facile one-pot hydrothermal synthesis of hematite (α-Fe2O3) nanostructures and cephalexin antibiotic sorptive removal from polluted aqueous media. J Mol Liq 271:844–856. https​://doi.org/10.1016/j.molli​q.2018.09.057 72. Wilson D, Langell MA (2014) XPS analysis of oleylamine/oleic acid capped ­Fe3O4 nanoparticles as a function of temperature. Appl Surf Sci 303:6–13. https​://doi.org/10.1016/j.apsus​c.2014.02.006 73. Brewster DA, Sarappa DJ, Knowles KE (2019) Role of aliphatic ligands and solvent composition in the solvothermal synthesis of iron oxide nanocrystals. Polyhedron 157:54–62. https​://doi. org/10.1016/j.poly.2018.09.063 74. Huang J, Han J, Wang R, Zhang Y, Wang X, Zhang X, Zhang Z, Zhang Y, Song B, Jin S (2018) Improving electrocatalysts for oxygen evolution using N ­ ixFe3–xO4/Ni hybrid nanostructures formed by solvothermal synthesis. ACS Energy Lett 3(7):1698–1707. https​://doi.org/10.1021/acsen​ergyl​ ett.8b008​88 75. Kim J, Tran VT, Oh S, Kim C-S, Hong JC, Kim S, Joo Y-S, Mun S, Kim M-H, Jung J-W, Lee J, Kang YS, Koo J-W, Lee J (2018) Scalable solvothermal synthesis of superparamagnetic F ­ e3O4 nanoclusters for bioseparation and theragnostic probes. ACS Appl Mater Interfaces 10(49):41935– 41946. https​://doi.org/10.1021/acsam​i.8b141​56 76. Xiao J, Zhang G, Qian J, Sun X, Tian J, Zhong K, Cai D, Wu Z (2018) Fabricating high-performance T2-weighted contrast agents via adjusting composition and size of nanomagnetic iron oxide. ACS Appl Mater Interfaces 10(8):7003–7011. https​://doi.org/10.1021/acsam​i.8b004​28 77. Köçkar H, Karaagac O, Özel F (2019) Effects of biocompatible surfactants on structural and corresponding magnetic properties of iron oxide nanoparticles coated by hydrothermal process. J Magn Magn Mater 474:332–336. https​://doi.org/10.1016/j.jmmm.2018.11.053 78. Fievet F, Lagier JP, Blin B, Beaudoin B, Figlarz M (1989) Homogeneous and heterogeneous nucleations in the polyol process for the preparation of micron and submicron size metal particles. Solid State Ion 32–33:198–205. https​://doi.org/10.1016/0167-2738(89)90222​-1 79. Viau G, Fiévet-Vincent F, Fiévet F, Toneguzzo P, Ravel F, Acher O (1997) Size dependence of microwave permeability of spherical ferromagnetic particles. J Appl Phys 81(6):2749–2754. https​ ://doi.org/10.1063/1.36397​9 80. Chakroune N, Viau G, Ricolleau C, Fiévet-Vincent F, Fiévet F (2003) Cobalt-based anisotropic particles prepared by the polyol process. J Mater Chem 13(2):312–318. https​://doi.org/10.1039/ B2093​83A 81. Viau G, Toneguzzo P, Pierrard A, Acher O, Fiévet-Vincent F, Fiévet F (2001) Heterogeneous nucleation and growth of metal nanoparticles in polyols. Script Mater 44(8):2263–2267. https​:// doi.org/10.1016/S1359​-6462(01)00752​-7 82. Ammar S, Helfen A, Jouini N, Fiévet F, Rosenman I, Villain F, Molinié P, Danot M (2001) Magnetic properties of ultrafine cobalt ferrite particles synthesized by hydrolysis in a polyol medium. J Mater Chem 11(1):186–192. https​://doi.org/10.1039/B0031​93N 83. Chow GM, Kurihara LK, Kemner KM, Schoen PE, Elam WT, Ervin A, Keller S, Zhang YD, Budnick J, Ambrose T (2011) Structural, morphological, and magnetic study of nanocrystalline cobalt-copper powders synthesized by the polyol process. J Mater Res 10(6):1546–1554. https​:// doi.org/10.1557/JMR.1995.1546 84. Jungk HO, Feldmann C (2011) Nonagglomerated, submicron α-Fe2O3 particles: preparation and application. J Mater Res 15(10):2244–2248. https​://doi.org/10.1557/JMR.2000.0322 85. Caruntu D, Caruntu G, Chen Y, O’Connor CJ, Goloverda G, Kolesnichenko VL (2004) Synthesis of variable-sized nanocrystals of ­Fe3O4 with high surface reactivity. Chem Mater 16(25):5527– 5534. https​://doi.org/10.1021/cm048​7977

13

78

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



86. Joseyphus RJ, Kodama D, Matsumoto T, Sato Y, Jeyadevan B, Tohji K (2007) Role of polyol in the synthesis of Fe particles. J Magn Magn Mater 310(2 Part 3):2393–2395. https​://doi.org/10.1016/j. jmmm.2006.10.1132 87. Cheng C, Xu F, Gu H (2011) Facile synthesis and morphology evolution of magnetic iron oxide nanoparticles in different polyol processes. New J Chem 35(5):1072–1079. https​://doi.org/10.1039/ C0NJ0​0986E​ 88. Hachani R, Lowdell M, Birchall M, Hervault A, Mertz D, Begin-Colin S, Thanh NTK (2016) Polyol synthesis, functionalisation, and biocompatibility studies of superparamagnetic iron oxide nanoparticles as potential MRI contrast agents. Nanoscale 8(6):3278–3287. https​://doi.org/10.1039/ C5NR0​3867G​ 89. Gavilán H, Sánchez EH, Brollo MEF, Asín L, Moerner KK, Frandsen C, Lázaro FJ, Serna CJ, Veintemillas-Verdaguer S, Morales MP, Gutiérrez L (2017) Formation mechanism of maghemite nanoflowers synthesized by a polyol-mediated process. ACS Omega 2(10):7172–7184. https​://doi. org/10.1021/acsom​ega.7b009​75 90. Hu F, MacRenaris KW, Waters EA, Liang T, Schultz-Sikma EA, Eckermann AL, Meade TJ (2009) Ultrasmall, water-soluble magnetite nanoparticles with high relaxivity for magnetic resonance imaging. J Phys Chem C 113(49):20855–20860. https​://doi.org/10.1021/jp907​216g 91. Wan J, Cai W, Meng X, Liu E (2007) Monodisperse water-soluble magnetite nanoparticles prepared by polyol process for high-performance magnetic resonance imaging. Chem Commun 47:5004–5006. https​://doi.org/10.1039/B7127​95B 92. Cai W, Wan J (2007) Facile synthesis of superparamagnetic magnetite nanoparticles in liquid polyols. J Colloid Interface Sci 305(2):366–370. https​://doi.org/10.1016/j.jcis.2006.10.023 93. Hemery G, Keyes AC, Garaio E, Rodrigo I, Garcia JA, Plazaola F, Garanger E, Sandre O (2017) Tuning sizes, morphologies, and magnetic properties of monocore versus multicore iron oxide nanoparticles through the controlled addition of water in the polyol synthesis. Inorg Chem 56(14):8232–8243. https​://doi.org/10.1021/acs.inorg​chem.7b009​56 94. Wetegrove M, Witte K, Bodnar W, Pfahl D-E, Springer A, Schell N, Westphal F, Burkel E (2019) Formation of maghemite nanostructures in polyol: tuning the particle size via the precursor stoichiometry. CrystEngComm 21(12):1956–1966. https​://doi.org/10.1039/C8CE0​2115E​ 95. Maity D, Chandrasekharan P, Si-Shen F, Xue J-M, Ding J (2010) Polyol-based synthesis of hydrophilic magnetite nanoparticles. J Appl Phys 107(9):09B310. https​://doi.org/10.1063/1.33558​98 96. Qu H, Ma H, Riviere A, Zhou W, O’Connor CJ (2012) One-pot synthesis in polyamines for preparation of water-soluble magnetite nanoparticles with amine surface reactivity. J Mater Chem 22(8):3311–3313. https​://doi.org/10.1039/C2JM1​5932E​ 97. Wang J, Zhang B, Wang L, Wang M, Gao F (2015) One-pot synthesis of water-soluble superparamagnetic iron oxide nanoparticles and their MRI contrast effects in the mouse brains. Mater Sci Eng, C 48:416–423. https​://doi.org/10.1016/j.msec.2014.12.026 98. Kandasamy G, Soni S, Sushmita K, Veerapu NS, Bose S, Maity D (2019) One-step synthesis of hydrophilic functionalized and cytocompatible superparamagnetic iron oxide nanoparticles (SPIONs) based aqueous ferrofluids for biomedical applications. J Mol Liq 274:653–663. https​://doi. org/10.1016/j.molli​q.2018.10.161 99. Miao C, Hu F, Rui Y, Duan Y, Gu H (2019) A T1/T2 dual functional iron oxide MRI contrast agent with super stability and low hypersensitivity benefited by ultrahigh carboxyl group density. J Mater Chem B 7(12):2081–2091. https​://doi.org/10.1039/C9TB0​0002J​ 100. Babić-Stojić B, Jokanović V, Milivojević D, Požek M, Jagličić Z, Makovec D, Orsini NJ, Marković M, Arsikin K, Paunović V (2018) Ultrasmall iron oxide nanoparticles: magnetic and NMR relaxometric properties. Curr Appl Phys 18(2):141–149. https​://doi.org/10.1016/j.cap.2017.11.017 101. Yoo D, Lee C, Seo B, Piao Y (2017) One pot synthesis of amine-functionalized and angular-shaped superparamagnetic iron oxide nanoparticles for MR/fluorescence bimodal imaging application. RSC Adv 7(21):12876–12885. https​://doi.org/10.1039/C6RA2​8495G​ 102. Wagle DV, Rondinone AJ, Woodward JD, Baker GA (2017) Polyol synthesis of magnetite nanocrystals in a thermostable ionic liquid. Cryst Growth Des 17(4):1558–1567. https​://doi. org/10.1021/acs.cgd.6b015​11 103. Durães L, Costa BFO, Vasques J, Campos J, Portugal A (2005) Phase investigation of as-prepared iron oxide/hydroxide produced by sol–gel synthesis. Mater Lett 59(7):859–863. https​:// doi.org/10.1016/j.matle​t.2004.10.066

Reprinted from the journal

79

13



Topics in Current Chemistry (2020) 378:40

104. Masthoff IC, Kraken M, Menzel D, Litterst FJ, Garnweitner G (2016) Study of the growth of hydrophilic iron oxide nanoparticles obtained via the non-aqueous sol–gel method. J Sol-Gel Sci Technol 77(3):553–564. https​://doi.org/10.1007/s1097​1-015-3883-1 105. Venturini J, Wermuth TB, Machado MC, Arcaro S, Alves AK, da Cas VA, Bergmann CP (2019) The influence of solvent composition in the sol–gel synthesis of cobalt ferrite (­CoFe2O4): a route to tuning its magnetic and mechanical properties. J Eur Ceram Soc 39(12):3442–3449. https​://doi.org/10.1016/j.jeurc​erams​oc.2019.01.030 106. Liu XQ, Tao SW, Shen YS (1997) Preparation and characterization of nanocrystalline α-Fe2O3 by a sol–gel process. Sens Actuators B Chem 40(2):161–165. https​://doi.org/10.1016/S0925​ -4005(97)80256​-0 107. Akbar A, Yousaf H, Riaz S, Naseem S (2019) Role of precursor to solvent ratio in tuning the magnetization of iron oxide thin films—a sol–gel approach. J Magn Magn Mater 471:14–24. https​://doi.org/10.1016/j.jmmm.2018.09.008 108. Ba-Abbad MM, Takriff MS, Benamor A, Mohammad AW (2017) Size and shape controlled of α-Fe2O3 nanoparticles prepared via sol–gel technique and their photocatalytic activity. J SolGel Sci Technol 81(3):880–893. https​://doi.org/10.1007/s1097​1-016-4228-4 109. Hu P, Chang T, Chen W-J, Deng J, Li S-L, Zuo Y-G, Kang L, Yang F, Hostetter M, Volinsky AA (2019) Temperature effects on magnetic properties of F ­ e3O4 nanoparticles synthesized by the sol–gel explosion-assisted method. J Alloy Compd 773:605–611. https​://doi.org/10.1016/j. jallc​om.2018.09.238 110. Danks AE, Hall SR, Schnepp Z (2016) The evolution of ‘sol–gel’ chemistry as a technique for materials synthesis. Mater Horizons 3(2):91–112. https​://doi.org/10.1039/C5MH0​0260E​ 111. Akbar A, Riaz S, Ashraf R, Naseem S (2015) Magnetic and magnetization properties of iron oxide thin films by microwave assisted sol–gel route. J Sol-Gel Sci Technol 74(2):320–328. https​://doi.org/10.1007/s1097​1-014-3528-9 112. Kopanja L, Milosevic I, Panjan M, Damnjanovic V, Tadic M (2016) Sol–gel combustion synthesis, particle shape analysis and magnetic properties of hematite (α-Fe2O3) nanoparticles embedded in an amorphous silica matrix. Appl Surf Sci 362:380–386. https​://doi.org/10.1016/j. apsus​c.2015.11.238 113. Kralj S, Makovec D (2015) Magnetic assembly of superparamagnetic iron oxide nanoparticle clusters into nanochains and nanobundles. ACS Nano 9(10):9700–9707. https​://doi. org/10.1021/acsna​no.5b023​28 114. Bagwe RP, Kanicky JR, Palla BJ, Patanjali PK, Shah DO (2001) Improved drug delivery using microemulsions: rationale, recent progress, and new horizons. Crit Rev Ther Drug Carrier Syst 18(1):77–140 115. Okoli C, Sanchez-Dominguez M, Boutonnet M, Jaras S, Civera C, Solans C, Kuttuva GR (2012) Comparison and functionalization study of microemulsion-prepared magnetic iron oxide nanoparticles. Langmuir 28(22):8479–8485. https​://doi.org/10.1021/la300​599q 116. Inouye K, Endo R, Otsuka Y, Miyashiro K, Kaneko K, Ishikawa T (1982) Oxygenation of ferrous ions in reversed micelle and reversed microemulsion. J Phys Chem 86(8):1465–1469. https​ ://doi.org/10.1021/j1003​97a05​1 117. Lawrence MJ (1994) Surfactant systems: microemulsions and vesicles as vehicles for drug delivery. Eur J Drug Metab Pharmacokinet 19(3):257–269. https​://doi.org/10.1007/BF031​ 88929​ 118. Lawrence MJ, Rees GD (2000) Microemulsion-based media as novel drug delivery systems. Adv Drug Deliv Rev 45(1):89–121. https​://doi.org/10.1016/S0169​-409X(00)00103​-4 119. Fendler JH (1987) Atomic and molecular clusters in membrane mimetic chemistry. Chem Rev 87(5):877–899. https​://doi.org/10.1021/cr000​81a00​2 120. Munshi N, De TK, Maitra A (1997) Size modulation of polymeric nanoparticles under controlled dynamics of microemulsion droplets. J Colloid Interface Sci 190(2):387–391. https​://doi. org/10.1006/jcis.1997.4889 121. Gupta AK, Wells S (2004) Surface-modified superparamagnetic nanoparticles for drug delivery: preparation, characterization, and cytotoxicity studies. IEEE Trans Nanobiosci 3(1):66–73. https​:// doi.org/10.1109/TNB.2003.82027​7 122. Igartua M, Saulnier P, Heurtault B, Pech B, Proust JE, Pedraz JL, Benoit JP (2002) Development and characterization of solid lipid nanoparticles loaded with magnetite. Int J Pharm 233(1):149– 157. https​://doi.org/10.1016/S0378​-5173(01)00936​-X

13

80

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



123. Yaacob II, Nunes AC, Bose A, Shah DO (1994) Synthesis and characterization of magnetic nanoparticles in spontaneously generated vesicles. J Colloid Interface Sci 168(2):289–301. https​://doi. org/10.1006/jcis.1994.1423 124. Uchida M, Flenniken ML, Allen M, Willits DA, Crowley BE, Brumfield S, Willis AF, Jackiw L, Jutila M, Young MJ, Douglas T (2006) Targeting of cancer cells with ferrimagnetic ferritin cage nanoparticles. J Am Chem Soc 128(51):16626–16633. https​://doi.org/10.1021/ja065​5690 125. Bhandarkar S, Bose A (1990) Synthesis of nanocomposite particles by intravesicular coprecipitation. J Colloid Interface Sci 139(2):541–550. https​://doi.org/10.1016/0021-9797(90)90127​-A 126. Yaacob II, Nunes AC, Bose A (1995) Magnetic nanoparticles produced in spontaneous cationic-anionic vesicles: room temperature synthesis and characterization. J Colloid Interface Sci 171(1):73–84. https​://doi.org/10.1006/jcis.1995.1152 127. Najafi A, Nematipour K (2017) Synthesis and magnetic properties evaluation of monosized FeCo alloy nanoparticles through microemulsion method. J Supercond Novel Magn 30(9):2647–2653. https​://doi.org/10.1007/s1094​8-017-4052-2 128. Singh P, Upadhyay C (2018) Fine tuning of size and morphology of magnetite nanoparticles synthesized by microemulsion. AIP Conf Proc 1953 1:030051. https​://doi.org/10.1063/1.50323​86 129. Bonacchi D, Caneschi A, Dorignac D, Falqui A, Gatteschi D, Rovai D, Sangregorio C, Sessoli R (2004) Nanosized iron oxide particles entrapped in pseudo-single crystals of γ-cyclodextrin. Chem Mater 16(10):2016–2020. https​://doi.org/10.1021/cm034​948e 130. Lee Y, Lee J, Bae CJ, Park JG, Noh HJ, Park JH, Hyeon T (2005) Large-scale synthesis of uniform and crystalline magnetite nanoparticles using reverse micelles as nanoreactors under reflux conditions. Adv Func Mater 15(3):503–509. https​://doi.org/10.1002/adfm.20040​0187 131. Vidal-Vidal J, Rivas J, López-Quintela MA (2006) Synthesis of monodisperse maghemite nanoparticles by the microemulsion method. Colloids Surf A 288(1):44–51. https​://doi.org/10.1016/j.colsu​ rfa.2006.04.027 132. Pileni M-P (2003) The role of soft colloidal templates in controlling the size and shape of inorganic nanocrystals. Nat Mater 2(3):145–150. https​://doi.org/10.1038/nmat8​17 133. Han X, Yao P, Cheng C, Yuan H, Yang Y, Ni C (2018) Preparation and in vivo biodistribution of ultra-small superparamagnetic iron oxide nanoparticles with high magnetic targeting response. J Nanosci Nanotechnol 18(2):879–886. https​://doi.org/10.1166/jnn.2018.14110​ 134. De Cuyper M, Joniau M (1988) Magnetoliposomes. Eur Biophys J 15(5):311–319. https​://doi. org/10.1007/BF002​56482​ 135. Rocha FM, de Pinho SC, Zollner RL, Santana MHA (2001) Preparation and characterization of affinity magnetoliposomes useful for the detection of antiphospholipid antibodies. J Magn Magn Mater 225(1):101–108. https​://doi.org/10.1016/S0304​-8853(00)01236​-1 136. Bulte JWM, Cuyper Md, Despres D, Frank JA (1999) Preparation, relaxometry, and biokinetics of PEGylated magnetoliposomes as MR contrast agent. J Magn Magn Mater 194(1):204–209. https​:// doi.org/10.1016/S0304​-8853(98)00556​-3 137. Lesieur S, Grabielle-Madelmont C, Ménager C, Cabuil V, Dadhi D, Pierrot P, Edwards K (2003) Evidence of surfactant-induced formation of transient pores in lipid bilayers by using magneticfluid-loaded liposomes. J Am Chem Soc 125(18):5266–5267. https​://doi.org/10.1021/ja021​471j 138. Nobuto H, Sugita T, Kubo T, Shimose S, Yasunaga Y, Murakami T, Ochi M (2004) Evaluation of systemic chemotherapy with magnetic liposomal doxorubicin and a dipole external electromagnet. Int J Cancer 109(4):627–635. https​://doi.org/10.1002/ijc.20035​ 139. Sangregorio C, Wiemann JK, O’Connor CJ, Rosenzweig Z (1999) A new method for the synthesis of magnetoliposomes. J Appl Phys 85(8):5699–5701. https​://doi.org/10.1063/1.37025​6 140. Kaur G, Dogra V, Kumar R, Kumar S, Singh K (2019) Fabrication of iron oxide nanocolloids using metallosurfactant-based microemulsions: antioxidant activity, cellular, and genotoxicity toward Vitis vinifera. J Biomol Struct Dyn 37(4):892–909. https​://doi.org/10.1080/07391​102.2018.14422​ 51 141. Lee C, Kim GR, Yoon J, Kim SE, Yoo JS, Piao Y (2018) In  vivo delineation of glioblastoma by targeting tumor-associated macrophages with near-infrared fluorescent silica coated iron oxide nanoparticles in orthotopic xenografts for surgical guidance. Sci Rep 8(1):11122. https​://doi. org/10.1038/s4159​8-018-29424​-4 142. Kampferbeck M, Vossmeyer T, Weller H (2019) Cross-linked polystyrene shells grown on iron oxide nanoparticles via surface-grafted AGET–ATRP in microemulsion. Langmuir 35(26):8790– 8798. https​://doi.org/10.1021/acs.langm​uir.9b010​60

Reprinted from the journal

81

13



Topics in Current Chemistry (2020) 378:40

143. Cao Y, Min J, Zheng D, Li J, Xue Y, Yu F, Wu M (2019) Vehicle-saving theranostic probes based on hydrophobic iron oxide nanoclusters using doxorubicin as a phase transfer agent for MRI and chemotherapy. Chem Commun 55(61):9015–9018. https​://doi.org/10.1039/C9CC0​3868J​ 144. Rodríguez-Rodríguez AA, Moreno-Trejo MB, Meléndez-Zaragoza MJ, Collins-Martínez V, LópezOrtiz A, Martínez-Guerra E, Sánchez-Domínguez M (2019) Spinel-type ferrite nanoparticles: synthesis by the oil-in-water microemulsion reaction method and photocatalytic water-splitting evaluation. Int J Hydrogen Energy 44(24):12421–12429. https​://doi.org/10.1016/j.ijhyd​ene.2018.09.183 145. Amirshaghaghi A, Yan L, Miller J, Daniel Y, Stein JM, Busch TM, Cheng Z, Tsourkas A (2019) Chlorin e6-coated superparamagnetic iron oxide nanoparticle (SPION) nanoclusters as a theranostic agent for dual-mode imaging and photodynamic therapy. Sci Rep 9(1):2613. https​://doi. org/10.1038/s4159​8-019-39036​-1 146. Pecharromán C, González-Carreño T, Iglesias JE (1995) The infrared dielectric properties of maghemite, γ-Fe2O3, from reflectance measurement on pressed powders. Phys Chem Miner 22(1):21–29. https​://doi.org/10.1007/BF002​02677​ 147. González-Carreño T, Morales MP, Gracia M, Serna CJ (1993) Preparation of uniform γ-Fe2O3 particles with nanometer size by spray pyrolysis. Mater Lett 18(3):151–155. https​://doi. org/10.1016/0167-577X(93)90116​-F 148. Che S, Sakurai O, Shinozaki K, Mizutani N (1998) Particle structure control through intraparticle reactions by spray pyrolysis. J Aerosol Sci 29(3):271–278. https​://doi.org/10.1016/S0021​ -8502(97)10012​-X 149. Kastrinaki G, Lorentzou S, Karagiannakis G, Rattenbury M, Woodhead J, Konstandopoulos AG (2018) Parametric synthesis study of iron based nanoparticles via aerosol spray pyrolysis route. J Aerosol Sci 115:96–107. https​://doi.org/10.1016/j.jaero​sci.2017.10.005 150. Zheng J, Liu K, Cai W, Qiao L, Ying Y, Li W, Yu J, Lin M, Che S (2018) Effect of chloride ion on crystalline phase transition of iron oxide produced by ultrasonic spray pyrolysis. Adv Powder Technol 29(9):1953–1959. https​://doi.org/10.1016/j.apt.2018.03.028 151. Das H, Debnath N, Toda A, Kawaguchi T, Sakamoto N, Manjura Hoque S, Shinozaki K, Suzuki H, Wakiya N (2018) Controlled synthesis of dense ­MgFe2O4 nanospheres by ultrasonic spray pyrolysis technique: effect of ethanol addition to precursor solvent. Adv Powder Technol 29(2):283–288. https​://doi.org/10.1016/j.apt.2017.11.014 152. Morales MP, Bomati-Miguel O, Pérez de Alejo R, Ruiz-Cabello J, Veintemillas-Verdaguer S, O’Grady K (2003) Contrast agents for MRI based on iron oxide nanoparticles prepared by laser pyrolysis. J Magn Magn Mater 266(1):102–109. https​://doi.org/10.1016/S0304​-8853(03)00461​-X 153. Miguel OB, Morales MP, Serna CJ, Veintemillas-Verdaguer S (2002) Magnetic nanoparticles prepared by laser pyrolysis. IEEE Trans Magn 38(5):2616–2618. https​://doi.org/10.1109/ TMAG.2002.80196​1 154. Zhao XQ, Zheng F, Liang Y, Hu ZQ, Xu YB (1994) Preparation and characterization of single phase γ-Fe nanopowder from cw C ­ O2 laser induced pyrolysis of iron pentacarbonyl. Mater Lett 21(3):285–288. https​://doi.org/10.1016/0167-577X(94)90191​-0 155. Julián-López B, Boissière C, Chanéac C, Grosso D, Vasseur S, Miraux S, Duguet E, Sanchez C (2007) Mesoporous maghemite–organosilica microspheres: a promising route towards multifunctional platforms for smart diagnosis and therapy. J Mater Chem 17(16):1563–1569. https​://doi. org/10.1039/B6159​51F 156. Teoh WY, Amal R, Mädler L (2010) Flame spray pyrolysis: an enabling technology for nanoparticles design and fabrication. Nanoscale 2(8):1324–1347. https​://doi.org/10.1039/C0NR0​0017E​ 157. Mädler L, Kammler HK, Mueller R, Pratsinis SE (2002) Controlled synthesis of nanostructured particles by flame spray pyrolysis. J Aerosol Sci 33(2):369–389. https​://doi.org/10.1016/S0021​ -8502(01)00159​-8 158. Xu H, Zeiger BW, Suslick KS (2013) Sonochemical synthesis of nanomaterials. Chem Soc Rev 42(7):2555–2567. https​://doi.org/10.1039/C2CS3​5282F​ 159. Suslick KS, Nyborg WL (1990) Ultrasound: its chemical, physical and biological effects. J Acoust Soc Am 87(2):919–920. https​://doi.org/10.1121/1.39886​4 160. Suslick KS, Didenko Y, Fang MM, Hyeon T, Kolbeck KJ, McNamara WB, Mdleleni MM, Wong M (1999) Acoustic cavitation and its chemical consequences. Philos Trans R Soc Lond Ser A Math Phys Eng Sci 357(1751):335–353. https​://doi.org/10.1098/rsta.1999.0330 161. Dolores R, Raquel S, Adianez G-L (2015) Sonochemical synthesis of iron oxide nanoparticles loaded with folate and cisplatin: effect of ultrasonic frequency. Ultrason Sonochem 23:391–398. https​://doi.org/10.1016/j.ultso​nch.2014.08.005

13

82

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



162. Shafi KVPM, Ulman A, Yan X, Yang N-L, Estournès C, White H, Rafailovich M (2001) Sonochemical synthesis of functionalized amorphous iron oxide nanoparticles. Langmuir 17(16):5093– 5097. https​://doi.org/10.1021/la010​421+ 163. Vijayakumar R, Koltypin Y, Felner I, Gedanken A (2000) Sonochemical synthesis and characterization of pure nanometer-sized F ­ e3O4 particles. Mater Sci Eng A 286(1):101–105. https​://doi. org/10.1016/S0921​-5093(00)00647​-X 164. Abu Mukh-Qasem R, Gedanken A (2005) Sonochemical synthesis of stable hydrosol of F ­ e3O4 nanoparticles. J Colloid Interface Sci 284(2):489–494. https​://doi.org/10.1016/j.jcis.2004.10.073 165. Rahmawati R, Kaneti YV, Taufiq A, Sunaryono YB, Suyatman N, Kurniadi D, Hossain MSA, Yamauchi Y (2018) Green synthesis of magnetite nanostructures from naturally available iron sands via sonochemical method. Bull Chem Soc Jpn 91(2):311–317. https​://doi.org/10.1246/ bcsj.20170​317 166. Hee Kim E, Sook Lee H, Kook Kwak B, Kim B-K (2005) Synthesis of ferrofluid with magnetic nanoparticles by sonochemical method for MRI contrast agent. J Magn Magn Mater 289:328–330. https​://doi.org/10.1016/j.jmmm.2004.11.093 167. Abbas M, Torati SR, Kim C (2015) A novel approach for the synthesis of ultrathin silica-coated iron oxide nanocubes decorated with silver nanodots ­(Fe3O4/SiO2/Ag) and their superior catalytic reduction of 4-nitroaniline. Nanoscale 7(28):12192–12204. https​://doi.org/10.1039/C5NR0​2680F​ 168. Shafi KVPM, Koltypin Y, Gedanken A, Prozorov R, Balogh J, Lendvai J, Felner I (1997) Sonochemical preparation of nanosized amorphous N ­ iFe2O4 particles. J Phys Chem B 101(33):6409– 6414. https​://doi.org/10.1021/jp970​893q 169. Shafi KVPM, Gedanken A, Prozorov R, Balogh J (1998) Sonochemical preparation and sizedependent properties of nanostructured ­CoFe2O4 particles. Chem Mater 10(11):3445–3450. https​:// doi.org/10.1021/cm980​182k 170. Sodipo BK, Aziz AA (2018) One minute synthesis of amino-silane functionalized superparamagnetic iron oxide nanoparticles by sonochemical method. Ultrason Sonochem 40:837–840. https​:// doi.org/10.1016/j.ultso​nch.2017.08.040 171. Sathya A, Kalyani S, Ranoo S, Philip J (2017) One-step microwave-assisted synthesis of waterdispersible ­Fe3O4 magnetic nanoclusters for hyperthermia applications. J Magn Magn Mater 439:107–113. https​://doi.org/10.1016/j.jmmm.2017.05.018 172. Palchik O, Felner I, Kataby G, Gedanken A (2011) Amorphous iron oxide prepared by microwave heating. J Mater Res 15(10):2176–2181. https​://doi.org/10.1557/JMR.2000.0313 173. Liu Z, Miao F, Hua W, Zhao F (2012) ­Fe3O4 nanoparticles: microwave-assisted synthesis and mechanism. Mater Lett 67(1):358–361. https​://doi.org/10.1016/j.matle​t.2011.09.095 174. Kostyukhin EM, Kustov LM (2018) Microwave-assisted synthesis of magnetite nanoparticles possessing superior magnetic properties. Mendeleev Commun 28(5):559–561. https​://doi. org/10.1016/j.menco​m.2018.09.038 175. Bonfim L, de Queiroz Souza PP, de Oliveira Gonçalves K, Courrol LC, de Oliveira Silva FR, Vieira DP (2019) Microwave-mediated synthesis of iron-oxide nanoparticles for use in magnetic levitation cell cultures. Appl Nanosci 9(8):1707–1717. https​://doi.org/10.1007/s1320​4-019-00962​ -1 176. Aivazoglou E, Metaxa E, Hristoforou E (2017) Microwave-assisted synthesis of iron oxide nanoparticles in biocompatible organic environment. AIP Adv 8(4):048201. https​://doi. org/10.1063/1.49940​57 177. Lastovina TA, Budnyk AP, Soldatov MA, Rusalev YV, Guda AA, Bogdan AS, Soldatov AV (2017) Microwave-assisted synthesis of magnetic iron oxide nanoparticles in oleylamine–oleic acid solutions. Mendeleev Commun 27(5):487–489. https​://doi.org/10.1016/j.menco​m.2017.09.019 178. Cao S-W, Zhu Y-J (2009) Iron oxide hollow spheres: microwave–hydrothermal ionic liquid preparation, formation mechanism, crystal phase and morphology control and properties. Acta Mater 57(7):2154–2165. https​://doi.org/10.1016/j.actam​at.2009.01.009 179. Blanco-Andujar C, Ortega D, Southern P, Pankhurst QA, Thanh NTK (2015) High performance multi-core iron oxide nanoparticles for magnetic hyperthermia: microwave synthesis, and the role of core-to-core interactions. Nanoscale 7(5):1768–1775. https​://doi.org/10.1039/C4NR0​6239F​ 180. Jiang FY, Wang CM, Fu Y, Liu RC (2010) Synthesis of iron oxide nanocubes via microwaveassisted solvolthermal method. J Alloy Compd 503(2):L31–L33. https​://doi.org/10.1016/j.jallc​ om.2010.05.020

Reprinted from the journal

83

13

Topics in Current Chemistry (2020) 378:40 181. Hu L, Percheron A, Chaumont D, Brachais C-H (2011) Microwave-assisted one-step hydrothermal synthesis of pure iron oxide nanoparticles: magnetite, maghemite and hematite. J Sol-Gel Sci Technol 60(2):198. https​://doi.org/10.1007/s1097​1-011-2579-4 182. Wu L, Yao H, Hu B, Yu S-H (2011) Unique lamellar sodium/potassium iron oxide nanosheets: facile microwave-assisted synthesis and magnetic and electrochemical properties. Chem Mater 23(17):3946–3952. https​://doi.org/10.1021/cm201​3736 183. Katsuki H, Choi E-K, Lee W-J, Hwang K-T, Cho W-S, Huang W, Komarneni S (2018) Ultrafast microwave-hydrothermal synthesis of hexagonal plates of hematite. Mater Chem Phys 205:210– 216. https​://doi.org/10.1016/j.match​emphy​s.2017.10.078 184. Fernández-Barahona I, Gutiérrez L, Veintemillas-Verdaguer S, Pellico J, Morales MdP, Catala M, del Pozo MA, Ruiz-Cabello J, Herranz F (2019) Cu-doped extremely small iron oxide nanoparticles with large longitudinal relaxivity: one-pot synthesis and in vivo targeted molecular imaging. ACS Omega 4(2):2719–2727. https​://doi.org/10.1021/acsom​ega.8b030​04 185. Hong RY, Pan TT, Li HZ (2006) Microwave synthesis of magnetic ­Fe3O4 nanoparticles used as a precursor of nanocomposites and ferrofluids. J Magn Magn Mater 303(1):60–68. https​://doi. org/10.1016/j.jmmm.2005.10.230 186. Katsuki H, Komarneni S (2001) Microwave-hydrothermal synthesis of monodispersed nanophase α-Fe2O3. J Am Ceram Soc 84(10):2313–2317. https​://doi.org/10.1111/j.1151-2916.2001.tb010​07.x 187. Pascu O, Carenza E, Gich M, Estradé S, Peiró F, Herranz G, Roig A (2012) Surface reactivity of iron oxide nanoparticles by microwave-assisted synthesis; comparison with the thermal decomposition route. J Phys Chem C 116(28):15108–15116. https​://doi.org/10.1021/jp303​204d 188. Xiao W, Gu H, Li D, Chen D, Deng X, Jiao Z, Lin J (2012) Microwave-assisted synthesis of magnetite nanoparticles for MR blood pool contrast agents. J Magn Magn Mater 324(4):488–494. https​ ://doi.org/10.1016/j.jmmm.2011.08.029 189. Carenza E, Barceló V, Morancho A, Montaner J, Rosell A, Roig A (2014) Rapid synthesis of waterdispersible superparamagnetic iron oxide nanoparticles by a microwave-assisted route for safe labeling of endothelial progenitor cells. Acta Biomater 10(8):3775–3785. https​://doi.org/10.1016/j. actbi​o.2014.04.010 190. Narayanan KB, Sakthivel N (2010) Biological synthesis of metal nanoparticles by microbes. Adv Coll Interface Sci 156(1):1–13. https​://doi.org/10.1016/j.cis.2010.02.001 191. Shah M, Fawcett D, Sharma S, Tripathy KS, Poinern EG (2015) Green synthesis of metallic nanoparticles via biological entities. Materials 8:11. https​://doi.org/10.3390/ma811​5377 192. Lovley DR, Stolz JF, Nord GL, Phillips EJP (1987) Anaerobic production of magnetite by a dissimilatory iron-reducing microorganism. Nature 330(6145):252–254. https​://doi.org/10.1038/33025​ 2a0 193. Zhang C, Vali H, Romanek CS, Phelps TJ, Liu SV (1998) Formation of single-domain magnetite by a thermophilic bacterium. Am Mineral 83(11–12_Part_2):1409–1418. https​://doi.org/10.2138/ am-1998-11-1230 194. Philipse AP, Maas D (2002) Magnetic colloids from magnetotactic bacteria: chain formation and colloidal stability. Langmuir 18(25):9977–9984. https​://doi.org/10.1021/la020​5811 195. Lee H, Purdon AM, Chu V, Westervelt RM (2004) Controlled assembly of magnetic nanoparticles from magnetotactic bacteria using microelectromagnets arrays. Nano Lett 4(5):995–998. https​:// doi.org/10.1021/nl049​562x 196. Lang C, Schüler D (2006) Biogenic nanoparticles: production, characterization, and application of bacterial magnetosomes. J Phys Condens Matter 18(38):S2815–S2828. https​://doi. org/10.1088/0953-8984/18/38/s19 197. Lisy MR, Hartung A, Lang C, Schuler D, Richter W, Reichenbach JR, Kaiser WA, Hilger I (2007) Fluorescent bacterial magnetic nanoparticles as bimodal contrast agents. Invest Radiol 42(4):235– 241. https​://doi.org/10.1097/01.rli.00002​55832​.44443​.e7 198. Bharde A, Wani A, Shouche Y, Joy PA, Prasad BLV, Sastry M (2005) Bacterial aerobic synthesis of nanocrystalline magnetite. J Am Chem Soc 127(26):9326–9327. https​://doi.org/10.1021/ja050​ 8469 199. Bharde AA, Parikh RY, Baidakova M, Jouen S, Hannoyer B, Enoki T, Prasad BLV, Shouche YS, Ogale S, Sastry M (2008) Bacteria-mediated precursor-dependent biosynthesis of superparamagnetic iron oxide and iron sulfide nanoparticles. Langmuir 24(11):5787–5794. https​://doi. org/10.1021/la704​019p

13

84

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40



200. Bharde A, Rautaray D, Bansal V, Ahmad A, Sarkar I, Yusuf SM, Sanyal M, Sastry M (2006) Extracellular biosynthesis of magnetite using fungi. Small 2(1):135–141. https​://doi.org/10.1002/ smll.20050​0180 201. Shenton W, Douglas T, Young M, Stubbs G, Mann S (1999) Inorganic–organic nanotube composites from template mineralization of tobacco mosaic virus. Adv Mater 11(3):253–256. https​://doi. org/10.1002/(SICI)1521-4095(19990​3)11:3%3c253​:AID-ADMA2​53%3e3.0.CO;2-7 202. Aeppli M, Kaegi R, Kretzschmar R, Voegelin A, Hofstetter TB, Sander M (2019) Electrochemical analysis of changes in iron oxide reducibility during abiotic ferrihydrite transformation into goethite and magnetite. Environ Sci Technol 53(7):3568–3578. https​://doi.org/10.1021/acs.est.8b071​ 90 203. Moon J-W, Rawn CJ, Rondinone AJ, Love LJ, Roh Y, Everett SM, Lauf RJ, Phelps TJ (2010) Large-scale production of magnetic nanoparticles using bacterial fermentation. J Ind Microbiol Biotechnol 37(10):1023–1031. https​://doi.org/10.1007/s1029​5-010-0749-y 204. Iravani S (2011) Green synthesis of metal nanoparticles using plants. Green Chem 13(10):2638– 2650. https​://doi.org/10.1039/C1GC1​5386B​ 205. Aksu Demirezen D, Yıldız YŞ, Yılmaz Ş, Demirezen Yılmaz D (2019) Green synthesis and characterization of iron oxide nanoparticles using Ficus carica (common fig) dried fruit extract. J Biosci Bioeng 127(2):241–245. https​://doi.org/10.1016/j.jbios​c.2018.07.024 206. Lohrasbi S, Kouhbanani MAJ, Beheshtkhoo N, Ghasemi Y, Amani AM, Taghizadeh S (2019) Green synthesis of iron nanoparticles using plantago major leaf extract and their application as a catalyst for the decolorization of azo dye. BioNanoScience 9(2):317–322. https​://doi.org/10.1007/ s1266​8-019-0596-x 207. Salazar-Alvarez G, Muhammed M, Zagorodni AA (2006) Novel flow injection synthesis of iron oxide nanoparticles with narrow size distribution. Chem Eng Sci 61(14):4625–4633. https​://doi. org/10.1016/j.ces.2006.02.032 208. Dierstein A, Natter H, Meyer F, Stephan HO, Kropf C, Hempelmann R (2001) Electrochemical deposition under oxidizing conditions (EDOC): a new synthesis for nanocrystalline metal oxides. Scripta Mater 44(8):2209–2212. https​://doi.org/10.1016/S1359​-6462(01)00906​-X 209. Pascal C, Pascal JL, Favier F, Elidrissi Moubtassim ML, Payen C (1999) Electrochemical synthesis for the control of γ-Fe2O3 nanoparticle size. Morphology, microstructure, and magnetic behavior. Chem Mater 11(1):141–147. https​://doi.org/10.1021/cm980​742f 210. Ramimoghadam D, Bagheri S, Hamid SBA (2014) Progress in electrochemical synthesis of magnetic iron oxide nanoparticles. J Magn Magn Mater 368:207–229. https​://doi.org/10.1016/j. jmmm.2014.05.015 211. Carraro G, Barreca D, Maccato C, Bontempi E, Depero LE, de Julián FC, Caneschi A (2013) Supported ε and β iron oxide nanomaterials by chemical vapor deposition: structure, morphology and magnetic properties. CrystEngComm 15(6):1039–1042. https​://doi.org/10.1039/C2CE2​6821C​ 212. Alijani H, Beyki MH, Shariatinia Z, Bayat M, Shemirani F (2014) A new approach for one step synthesis of magnetic carbon nanotubes/diatomite earth composite by chemical vapor deposition method: application for removal of lead ions. Chem Eng J 253:456–463. https​://doi.org/10.1016/j. cej.2014.05.021 213. Morjan I, Alexandrescu R, Dumitrache F, Birjega R, Fleaca C, Soare I, Luculescu CR, Filoti G, Kuncer V, Vekas L, Popa NC, Prodan G, Ciupina V (2010) Iron oxide-based nanoparticles with different mean sizes obtained by the laser pyrolysis: structural and magnetic properties. J Nanosci Nanotechnol 10(2):1223–1234. https​://doi.org/10.1166/jnn.2010.1863 214. Dinesha ML, Jayanna HS, Mohanty S, Ravi S (2010) Structural, electrical and magnetic properties of Co and Fe co-doped ZnO nanoparticles prepared by solution combustion method. J Alloy Compd 490(1):618–623. https​://doi.org/10.1016/j.jallc​om.2009.10.120 215. Ma J, Lee SM-Y, Yi C, Li C-W (2017) Controllable synthesis of functional nanoparticles by microfluidic platforms for biomedical applications—a review. Lab Chip 17(2):209–226. https​://doi. org/10.1039/C6LC0​1049K​ 216. Hwang DK, Dendukuri D, Doyle PS (2008) Microfluidic-based synthesis of non-spherical magnetic hydrogel microparticles. Lab Chip 8(10):1640–1647. https​://doi.org/10.1039/B8051​76C 217. Wei J, Shuai X, Wang R, He X, Li Y, Ding M, Li J, Tan H, Fu Q (2017) Clickable and imageable multiblock polymer micelles with magnetically guided and PEG-switched targeting and release property for precise tumor theranosis. Biomaterials 145:138–153. https​://doi.org/10.1016/j.bioma​ teria​ls.2017.08.005

Reprinted from the journal

85

13

Topics in Current Chemistry (2020) 378:40 218. Li H, Yan K, Shang Y, Shrestha L, Liao R, Liu F, Li P, Xu H, Xu Z, Chu PK (2015) Folate-bovine serum albumin functionalized polymeric micelles loaded with superparamagnetic iron oxide nanoparticles for tumor targeting and magnetic resonance imaging. Acta Biomater 15:117–126. https​:// doi.org/10.1016/j.actbi​o.2015.01.006 219. Starmans LWE, Moonen RPM, Aussems-Custers E, Daemen MJAP, Strijkers GJ, Nicolay K, Grüll H (2015) Evaluation of iron oxide nanoparticle micelles for magnetic particle imaging (MPI) of thrombosis. PLoS ONE 10:3. https​://doi.org/10.1371/journ​al.pone.01192​57 220. Xiao S, Castro R, Rodrigues J, Shi X, Tomá H (2014) PAMAM dendrimer/pDNA functionalizedmagnetic iron oxide nanoparticles for gene delivery. J Biomed Nanotechnol 11(8):1418–1430. https​://doi.org/10.1166/jbn.2015.2101 221. Upponi JR, Jerajani K, Nagesha DK, Kulkarni P, Sridhar S, Ferris C, Torchilin VP (2018) Polymeric micelles: theranostic co-delivery system for poorly water-soluble drugs and contrast agents. Biomaterials 170:26–36. https​://doi.org/10.1016/j.bioma​teria​ls.2018.03.054 222. Tsai CH, Tang YH, Chen HT, Yao YW, Chien TC, Kao CL (2018) A selective glucose sensor: the cooperative effect of monoboronic acid-modified poly(amidoamine) dendrimers. Chem Commun 54(36):4577–4580. https​://doi.org/10.1039/c8cc0​0914g​ 223. Babamiri B, Hallaj R, Salimi A (2018) Ultrasensitive electrochemiluminescence immunoassay for simultaneous determination of CA125 and CA15-3 tumor markers based on PAMAM-sulfanilic acid-Ru(bpy)32+ and PAMAM-CdTe@CdS nanocomposite. Biosens Bioelectron 99:353–360. https​://doi.org/10.1016/j.bios.2017.07.062 224. Wang G, Fu L, Walker A, Chen X, Lovejoy DB, Hao M, Lee A, Chung R, Rizos H, Irvine M, Zheng M, Liu X, Lu Y, Shi B (2019) Label-free fluorescent poly(amidoamine) dendrimer for traceable and controlled drug delivery. Biomacromol 20(5):2148–2158. https​://doi.org/10.1021/acs. bioma​c.9b004​94 225. Najafi F, Salami-Kalajahi M, Roghani-Mamaqani H, Kahaie-Khosrowshahi A (2019) Effect of grafting ratio of poly(propylene imine) dendrimer onto gold nanoparticles on the properties of colloidal hybrids, their DOX loading and release behavior and cytotoxicity. Colloids Surf B 178:500– 507. https​://doi.org/10.1016/j.colsu​rfb.2019.03.050 226. Wang B, Sun Y, Davis TP, Ke PC, Wu Y, Ding F (2018) Understanding effects of PAMAM dendrimer size and surface chemistry on serum protein binding with discrete molecular dynamics simulations. ACS Sustain Chem Eng 6(9):11704–11715. https​://doi.org/10.1021/acssu​schem​ eng.8b019​59 227. Tian F, Lin X, Valle RP, Zuo YY, Gu N (2019) Poly(amidoamine) dendrimer as a respiratory nanocarrier: insights from experiments and molecular dynamics simulations. Langmuir 35(15):5364– 5371. https​://doi.org/10.1021/acs.langm​uir.9b004​34 228. Jędrzak A, Grześkowiak BF, Coy E, Wojnarowicz J, Szutkowski K, Jurga S, Jesionowski T, Mrówczyński R (2019) Dendrimer based theranostic nanostructures for combined chemo- and photothermal therapy of liver cancer cells in  vitro. Colloids Surf B 173:698–708. https​://doi. org/10.1016/j.colsu​rfb.2018.10.045 229. Luong D, Sau S, Kesharwani P, Iyer AK (2017) Polyvalent Folate–Dendrimer-coated iron oxide theranostic nanoparticles for simultaneous magnetic resonance imaging and precise cancer cell targeting. Biomacromol 18(4):1197–1209. https​://doi.org/10.1021/acs.bioma​c.6b018​85 230. Shirmardi Shaghasemi B, Virk MM, Reimhult E (2017) Optimization of magneto-thermally controlled release kinetics by tuning of magnetoliposome composition and structure. Sci Rep 7:1. https​ ://doi.org/10.1038/s4159​8-017-06980​-9 231. German SV, Navolokin NA, Kuznetsova NR, Zuev VV, Inozemtseva OA, Anis’kov AA, Volkova EK, Bucharskaya AB, Maslyakova GN, Fakhrullin RF, Terentyuk GS, Vodovozova EL, Gorin DA (2015) Liposomes loaded with hydrophilic magnetite nanoparticles: preparation and application as contrast agents for magnetic resonance imaging. Colloids SurfB 135:109–115. https​://doi. org/10.1016/j.colsu​rfb.2015.07.042 232. Cuomo F, Cofelice M, Venditti F, Ceglie A, Miguel M, Lindman B, Lopez F (2018) In-vitro digestion of curcumin loaded chitosan-coated liposomes. Colloids Surf B 168:29–34. https​:// doi.org/10.1016/j.colsu​r fb.2017.11.047 233. Di Corato R, Béalle G, Kolosnjaj-Tabi J, Espinosa A, Clément O, Silva AKA, Ménager C, Wilhelm C (2015) Combining magnetic hyperthermia and photodynamic therapy for tumor ablation with photoresponsive magnetic liposomes. ACS Nano 9(3):2904–2916. https​://doi.org/10.1021/ nn506​949t

13

86

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40 234. Zheng XC, Ren W, Zhang S, Zhong T, Duan XC, Yin YF, Xu MQ, Hao YL, Li ZT, Li H, Liu M, Li ZY, Zhang X (2018) The theranostic efficiency of tumor-specific, pH-responsive, peptidemodified, liposome-containing paclitaxel and superparamagnetic iron oxide nanoparticles. Int J Nanomed 13:1495–1504. https​://doi.org/10.2147/IJN.S1570​82 235. Caro C, García-Martín ML, Pernia Leal M (2017) Manganese-based nanogels as pH switches for magnetic resonance imaging. Biomacromol 18(5):1617–1623. https​://doi.org/10.1021/acs. bioma​c.7b002​24 236. Riedinger A, Pernia Leal M, Deka SR, George C, Franchini IR, Falqui A, Cingolani R, Pellegrino T (2011) "nanohybrids" based on pH-responsive hydrogels and inorganic nanoparticles for drug delivery and sensor applications. Nano Lett 11(8):3136–3141 237. Jalili NA, Jaiswal MK, Peak CW, Cross LM, Gaharwar AK (2017) Injectable nanoengineered stimuli-responsive hydrogels for on-demand and localized therapeutic delivery. Nanoscale 9(40):15379–15389. https​://doi.org/10.1039/c7nr0​2327h​ 238. Peng N, Ding X, Wang Z, Cheng Y, Gong Z, Xu X, Gao X, Cai Q, Huang S, Liu Y (2019) Novel dual responsive alginate-based magnetic nanogels for onco-theranostics. Carbohyd Polym 204:32–41. https​://doi.org/10.1016/j.carbp​ol.2018.09.084 239. Li W, Xue B, Shi K, Qu Y, Chu B, Qian Z (2019) Magnetic iron oxide nanoparticles/10-hydroxy camptothecin co-loaded nanogel for enhanced photothermal-chemo therapy. Appl Mater Today 14:84–95. https​://doi.org/10.1016/j.apmt.2018.11.008 240. Sun W, Yang J, Zhu J, Zhou Y, Li J, Zhu X, Shen M, Zhang G, Shi X (2016) Immobilization of iron oxide nanoparticles within alginate nanogels for enhanced MR imaging applications. Biomater Sci 4(10):1422–1430. https​://doi.org/10.1039/c6bm0​0370b​ 241. Hao X, Xu B, Chen H, Wang X, Zhang J, Guo R, Shi X, Cao X (2019) Stem cell-mediated delivery of nanogels loaded with ultrasmall iron oxide nanoparticles for enhanced tumor MR imaging. Nanoscale 11(11):4904–4910. https​://doi.org/10.1039/c8nr1​0490e​ 242. Hwang J, Lee E, Kim J, Seo Y, Lee KH, Hong JW, Gilad AA, Park H, Choi J (2016) Effective delivery of immunosuppressive drug molecules by silica coated iron oxide nanoparticles. Colloids Surf B 142:290–296. https​://doi.org/10.1016/j.colsu​r fb.2016.01.040 243. Monaco I, Arena F, Biffi S, Locatelli E, Bortot B, La Cava F, Marini GM, Severini GM, Terreno E, Comes Franchini M (2017) Synthesis of lipophilic core-shell F ­ e3O4@SiO2@Au nanoparticles and polymeric entrapment into nanomicelles: a novel nanosystem for in vivo active targeting and magnetic resonance-photoacoustic dual imaging. Bioconjug Chem 28(5):1382–1390. https​://doi.org/10.1021/acs.bioco​njche​m.7b000​76 244. Guisasola E, Asín L, Beola L, de la Fuente JM, Baeza A, Vallet-Regí M (2018) Beyond traditional hyperthermia: in  vivo cancer treatment with magnetic-responsive mesoporous silica nanocarriers. ACS Appl Mater Interfaces 10(15):12518–12525. https​://doi.org/10.1021/acsam​ i.8b023​98 245. Hurley KR, Ring HL, Etheridge M, Zhang J, Gao Z, Shao Q, Klein ND, Szlag VM, Chung C, Reineke TM, Garwood M, Bischof JC, Haynes CL (2016) Predictable heating and positive MRI contrast from a mesoporous silica-coated iron oxide nanoparticle. Mol Pharm 13(7):2172–2183. https​://doi.org/10.1021/acs.molph​armac​eut.5b008​66 246. Park W, Yang HN, Ling D, Yim H, Kim KS, Hyeon T, Na K, Park KH (2014) Multi-modal transfection agent based on monodisperse magnetic nanoparticles for stem cell gene delivery and tracking. Biomaterials 35(25):7239–7247. https​://doi.org/10.1016/j.bioma​teria​ ls.2014.05.010 247. Lassenberger A, Bixner O, Gruenewald T, Lichtenegger H, Zirbs R, Reimhult E (2016) Evaluation of high-yield purification methods on monodisperse PEG-grafted iron oxide nanoparticles. Langmuir 32(17):4259–4269. https​://doi.org/10.1021/acs.langm​uir.6b009​19 248. Pernia Leal M, Rivera-Fernández S, Franco JM, Pozo D, de la Fuente JM, García-Martín ML (2015) Long-circulating PEGylated manganese ferrite nanoparticles for MRI-based molecular imaging. Nanoscale 7(5):2050–2059. https​://doi.org/10.1039/C4NR0​5781C​ 249. Pernia Leal M, Caro C, García-Martín ML (2017) Shedding light on zwitterionic magnetic nanoparticles: limitations for in vivo applications. Nanoscale 9(24):8176–8184. https​://doi.org/10.1039/ C7NR0​1607G​ 250. Leal MP, Muñoz-Hernández C, Berry CC, García-Martín ML (2015) In vivo pharmacokinetics of T2 contrast agents based on iron oxide nanoparticles: optimization of blood circulation times. RSC Adv 5(94):76883–76891. https​://doi.org/10.1039/C5RA1​5680G​

Reprinted from the journal

87

13



Topics in Current Chemistry (2020) 378:40

251. Weissleder R, Mahmood U (2001) Molecular imaging. Radiology 219(2):316–333. https​://doi. org/10.1148/radio​logy.219.2.r01ma​19316​ 252. Torrey HC (1956) Bloch equations with diffusion terms. Phys Rev 104(3):563–565. https​://doi. org/10.1103/PhysR​ev.104.563 253. Enriquez-Navas PM, Garcia-Martin ML (2012) Chapter 9—application of inorganic nanoparticles for diagnosis based on MRI. In: de la Fuente JM, Grazu V (eds) Frontiers of nanoscience, vol 4. Elsevier, Oxford, pp 233–245. https​://doi.org/10.1016/B978-0-12-41576​9-9.00009​-1 254. Geraldes CF, Laurent S (2009) Classification and basic properties of contrast agents for magnetic resonance imaging. Contrast Media Mol Imaging 4(1):1–23. https​://doi.org/10.1002/cmmi.265 255. Yan G-P, Robinson L, Hogg P (2007) Magnetic resonance imaging contrast agents: overview and perspectives. Radiography 13:e5–e19. https​://doi.org/10.1016/j.radi.2006.07.005 256. Bridot J-L, Faure A-C, Laurent S, Rivière C, Billotey C, Hiba B, Janier M, Josserand V, Coll J-L, Vander Elst L, Muller R, Roux S, Perriat P, Tillement O (2007) Hybrid gadolinium oxide nanoparticles: multimodal contrast agents for in vivo imaging. J Am Chem Soc 129(16):5076–5084. https​ ://doi.org/10.1021/ja068​356j 257. Ersoy H, Rybicki FJ (2007) Biochemical safety profiles of gadolinium-based extracellular contrast agents and nephrogenic systemic fibrosis. J Magn Reson Imaging 26(5):1190–1197. https​://doi. org/10.1002/jmri.21135​ 258. Dias MHM, Lauterbur PC (1986) Ferromagnetic particles as contrast agents for magnetic resonance imaging of liver and spleen. Magn Reson Med 3(2):328–330. https​://doi.org/10.1002/ mrm.19100​30218​ 259. Kim MP, Evans DB, Wang H, Abbruzzese JL, Fleming JB, Gallick GE (2009) Generation of orthotopic and heterotopic human pancreatic cancer xenografts in immunodeficient mice. Nat Protoc 4(11):1670–1680. https​://doi.org/10.1038/nprot​.2009.171 260. Jin K, Teng L, Shen Y, He K, Xu Z, Li G (2010) Patient-derived human tumour tissue xenografts in immunodeficient mice: a systematic review. Clin Transl Oncol 12(7):473–480. https​://doi. org/10.1007/s1209​4-010-0540-6 261. Carlos C, David P (2015) Polysaccharide colloids as smart vehicles in cancer therapy. Curr Pharm Des 21(33):4822–4836. https​://doi.org/10.2174/13816​12821​66615​08201​00812​ 262. Efremova MV, Naumenko VA, Spasova M, Garanina AS, Abakumov MA, Blokhina AD, Melnikov PA, Prelovskaya AO, Heidelmann M, Li Z-A, Ma Z, Shchetinin IV, Golovin YI, Kireev II, Savchenko AG, Chekhonin VP, Klyachko NL, Farle M, Majouga AG, Wiedwald U (2018) Magnetite–gold nanohybrids as ideal all-in-one platforms for theranostics. Sci Rep 8(1):11295. https​://doi. org/10.1038/s4159​8-018-29618​-w 263. Li Y, Song K, Cao Y, Peng C, Yang G (2018) Keratin-templated synthesis of metallic oxide nanoparticles as MRI contrast agents and drug carriers. ACS Appl Mater Interfaces 10(31):26039– 26045. https​://doi.org/10.1021/acsam​i.8b085​55 264. Hsu JC, Naha PC, Lau KC, Chhour P, Hastings R, Moon BF, Stein JM, Witschey WRT, McDonald ES, Maidment ADA, Cormode DP (2018) An all-in-one nanoparticle (AION) contrast agent for breast cancer screening with DEM-CT-MRI-NIRF imaging. Nanoscale 10(36):17236–17248. https​ ://doi.org/10.1039/C8NR0​3741H​ 265. Ren S, Yang J, Ma L, Li X, Wu W, Liu C, He J, Miao L (2018) Ternary-responsive drug delivery with activatable dual mode contrast-enhanced in  vivo imaging. ACS Appl Mater Interfaces 10(38):31947–31958. https​://doi.org/10.1021/acsam​i.8b105​64 266. Shirvalilou S, Khoei S, Khoee S, Raoufi NJ, Karimi MR, Shakeri-Zadeh A (2018) Development of a magnetic nano-graphene oxide carrier for improved glioma-targeted drug delivery and imaging: In  vitro and in  vivo evaluations. Chem Biol Interact 295:97–108. https​://doi.org/10.1016/j. cbi.2018.08.027 267. Sun J, Xu W, Li L, Fan B, Peng X, Qu B, Wang L, Li T, Li S, Zhang R (2018) Ultrasmall endogenous biopolymer nanoparticles for magnetic resonance/photoacoustic dual-modal imaging-guided photothermal therapy. Nanoscale 10(22):10584–10595. https​://doi.org/10.1039/C8NR0​1215F​ 268. Thirunavukkarasu GK, Cherukula K, Lee H, Jeong YY, Park I-K, Lee JY (2018) Magnetic fieldinducible drug-eluting nanoparticles for image-guided thermo-chemotherapy. Biomaterials 180:240–252. https​://doi.org/10.1016/j.bioma​teria​ls.2018.07.028 269. Zhang J, Mu Y-L, Ma Z-Y, Han K, Han H-Y (2018) Tumor-triggered transformation of chimeric peptide for dual-stage-amplified magnetic resonance imaging and precise photodynamic therapy. Biomaterials 182:269–278. https​://doi.org/10.1016/j.bioma​teria​ls.2018.08.026

13

88

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40 270. Sosnovik DE, Weissleder R (2007) Emerging concepts in molecular MRI. Curr Opin Biotechnol 18(1):4–10. https​://doi.org/10.1016/j.copbi​o.2006.11.001 271. Klausner RD (1996) The future of cancer research and the role of the National Cancer Institute. J Clin Oncol 14(10):2878–2883. https​://doi.org/10.1200/JCO.1996.14.10.2878 272. Gillies RJ (2002) In  vivo molecular imaging. J Cell Biochem 87(S39):231–238. https​://doi. org/10.1002/jcb.10450​ 273. Reimer P, Weissleder R, Lee AS, Wittenberg J, Brady TJ (1990) Receptor imaging: application to MR imaging of liver cancer. Radiology 177(3):729–734. https​://doi.org/10.1148/radio​ logy.177.3.22439​78 274. Tsoukalas C, Psimadas D, Kastis GA, Koutoulidis V, Harris AL, Paravatou-Petsotas M, Karageorgou M, Furenlid LR, Moulopoulos LA, Stamopoulos D, Bouziotis P (2018) A novel metalbased imaging probe for targeted dual-modality SPECT/MR imaging of angiogenesis. Front Chem 6:224–224. https​://doi.org/10.3389/fchem​.2018.00224​ 275. Ding N, Sano K, Kanazaki K, Ohashi M, Deguchi J, Kanada Y, Ono M, Saji H (2016) In  vivo HER2-targeted magnetic resonance tumor imaging using iron oxide nanoparticles conjugated with anti-HER2 fragment antibody. Mol Imag Biol 18(6):870–876. https​://doi.org/10.1007/s1130​ 7-016-0977-2 276. Ge Y, Zhong Y, Ji G, Lu Q, Dai X, Guo Z, Zhang P, Peng G, Zhang K, Li Y (2018) Preparation and characterization of ­Fe3O4@Au-C225 composite targeted nanoparticles for MRI of human glioma. PLoS ONE 13(4):e0195703–e0195703. https​://doi.org/10.1371/journ​al.pone.01957​03 277. Conde J, Bao C, Cui D, Baptista PV, Tian F (2014) Antibody–drug gold nanoantennas with Raman spectroscopic fingerprints for in  vivo tumour theranostics. J Control Release 183:87–93. https​:// doi.org/10.1016/j.jconr​el.2014.03.045 278. Zarschler K, Prapainop K, Mahon E, Rocks L, Bramini M, Kelly PM, Stephan H, Dawson KA (2014) Diagnostic nanoparticle targeting of the EGF-receptor in complex biological conditions using single-domain antibodies. Nanoscale 6(11):6046–6056. https​://doi.org/10.1039/C4NR0​ 0595C​ 279. Shevtsov MA, Nikolaev BP, Yakovleva LY, Marchenko YY, Dobrodumov AV, Mikhrina AL, Martynova MG, Bystrova OA, Yakovenko IV, Ischenko AM (2014) Superparamagnetic iron oxide nanoparticles conjugated with epidermal growth factor (SPION-EGF) for targeting brain tumors. Int J Nanomed 9:273–287. https​://doi.org/10.2147/IJN.S5511​8 280. Ren J, Zhang Z, Wang F, Yang Y, Liu Y, Wei G, Yang A, Zhang R, Huan Y, Cui Y, Larson AC (2012) MRI of prostate stem cell antigen expression in prostate tumors. Nanomedicine 7(5):691– 703. https​://doi.org/10.2217/nnm.11.147 281. Gottschalk K-E, Kessler H (2002) The structures of integrins and integrin–ligand complexes: implications for drug design and signal transduction. Angew Chem Int Ed 41(20):3767–3774. https​ ://doi.org/10.1002/1521-3773(20021​018)41:20%3c376​7:AID-ANIE3​767%3e3.0.CO;2-T 282. Akhtar MJ, Ahamed M, Alhadlaq HA, Alrokayan SA, Kumar S (2014) Targeted anticancer therapy: overexpressed receptors and nanotechnology. Clin Chim Acta 436:78–92. https​://doi. org/10.1016/j.cca.2014.05.004 283. Yameen B, Choi WI, Vilos C, Swami A, Shi J, Farokhzad OC (2014) Insight into nanoparticle cellular uptake and intracellular targeting. J Control Release 190:485–499. https​://doi.org/10.1016/j. jconr​el.2014.06.038 284. Chen K, Xie J, Xu H, Behera D, Michalski MH, Biswal S, Wang A, Chen X (2009) Triblock copolymer coated iron oxide nanoparticle conjugate for tumor integrin targeting. Biomaterials 30(36):6912–6919. https​://doi.org/10.1016/j.bioma​teria​ls.2009.08.045 285. Chen L, Xie J, Wu H, Zang F, Ma M, Hua Z, Gu N, Zhang Y (2018) Improving sensitivity of magnetic resonance imaging by using a dual-targeted magnetic iron oxide nanoprobe. Colloids Surf B Biointerfaces 161:339–346. https​://doi.org/10.1016/j.colsu​rfb.2017.10.059 286. Sánchez A, Ovejero Paredes K, Ruiz-Cabello J, Martínez-Ruíz P, Pingarrón JM, Villalonga R, Filice M (2018) Hybrid decorated core@shell janus nanoparticles as a flexible platform for targeted multimodal molecular bioimaging of cancer. ACS Appl Mater Interfaces 10(37):31032– 31043. https​://doi.org/10.1021/acsam​i.8b104​52 287. Li J, Wu C, Hou P, Zhang M, Xu K (2018) One-pot preparation of hydrophilic manganese oxide nanoparticles as T1 nano-contrast agent for molecular magnetic resonance imaging of renal carcinoma in  vitro and in  vivo. Biosens Bioelectron 102:1–8. https​://doi.org/10.1016/j. bios.2017.10.047

Reprinted from the journal

89

13



Topics in Current Chemistry (2020) 378:40

288. Zhao M, Liu Z, Dong L, Zhou H, Yang S, Wu W, Lin J (2018) A GPC3-specific aptamer-mediated magnetic resonance probe for hepatocellular carcinoma. Int J Nanomed 13:4433–4443. https​://doi.org/10.2147/IJN.S1682​68 289. Zhu Y, Sun Y, Chen Y, Liu W, Jiang J, Guan W, Zhang Z, Duan Y (2015) In  vivo molecular MRI imaging of prostate cancer by targeting PSMA with polypeptide-labeled superparamagnetic iron oxide nanoparticles. Int J Mol Sci 16(5):9573–9587. https​://doi.org/10.3390/ijms1​ 60595​73 290. Wang G, Qian K, Mei X (2018) A theranostic nanoplatform: magneto-gold@fluorescence polymer nanoparticles for tumor targeting T1&T2-MRI/CT/NIR fluorescence imaging and induction of genuine autophagy mediated chemotherapy. Nanoscale 10(22):10467–10478. https​://doi. org/10.1039/C8NR0​2429D​ 291. Jayapaul J, Arns S, Lederle W, Lammers T, Comba P, Gätjens J, Kiessling F (2012) Riboflavin carrier protein-targeted fluorescent USPIO for the assessment of vascular metabolism in tumors. Biomaterials 33(34):8822–8829. https​://doi.org/10.1016/j.bioma​teria​ls.2012.08.036 292. Chee HL, Gan CRR, Ng M, Low L, Fernig DG, Bhakoo KK, Paramelle D (2018) Biocompatible peptide-coated ultrasmall superparamagnetic iron oxide nanoparticles for in  vivo contrastenhanced magnetic resonance imaging. ACS Nano 12(7):6480–6491. https​://doi.org/10.1021/acsna​ no.7b075​72 293. Girard OM, Du J, Agemy L, Sugahara KN, Kotamraju VR, Ruoslahti E, Bydder GM, Mattrey RF (2011) Optimization of iron oxide nanoparticle detection using ultrashort echo time pulse sequences: comparison of T1, T2*, and synergistic T1–T2* contrast mechanisms. Magn Reson Med 65(6):1649–1660. https​://doi.org/10.1002/mrm.22755​ 294. Luo Y, Kim EH, Flask CA, Clark HA (2018) Nanosensors for the chemical imaging of acetylcholine using magnetic resonance imaging. ACS Nano 12(6):5761–5773. https​://doi.org/10.1021/ acsna​no.8b016​40 295. Bar-Shir A, Avram L, Yariv-Shoushan S, Anaby D, Cohen S, Segev-Amzaleg N, Frenkel D, Sadan O, Offen D, Cohen Y (2014) Alginate-coated magnetic nanoparticles for noninvasive MRI of extracellular calcium. NMR Biomed 27(7):774–783. https​://doi.org/10.1002/nbm.3117 296. Perez-Balderas F, van Kasteren SI, Aljabali AAA, Wals K, Serres S, Jefferson A, Sarmiento Soto M, Khrapitchev AA, Larkin JR, Bristow C, Lee SS, Bort G, De Simone F, Campbell SJ, Choudhury RP, Anthony DC, Sibson NR, Davis BG (2017) Covalent assembly of nanoparticles as a peptidasedegradable platform for molecular MRI. Nature Commun 8:14254. https​://doi.org/10.1038/ncomm​ s1425​4https​://www.natur​e.com/artic​les/ncomm​s1425​4#suppl​ement​ary-infor​matio​n 297. Henke PK, Pearce CG, Moaveni DM, Moore AJ, Lynch EM, Longo C, Varma M, Dewyer NA, Deatrick KB, Upchurch GR, Wakefield TW, Hogaboam C, Kunkel SL (2006) Targeted deletion of CCR2 impairs deep vein thombosis resolution in a mouse model. J Immunol 177(5):3388. https​:// doi.org/10.4049/jimmu​nol.177.5.3388 298. Yokoyama S, Ikeda H, Haramaki N, Yasukawa H, Murohara T, Imaizumi T (2005) Platelet P-selectin plays an important role in arterial thrombogenesis by forming large stable platelet-leukocyte aggregates. J Am Coll Cardiol 45(8):1280–1286. https​://doi.org/10.1016/j.jacc.2004.12.071 299. Suzuki M, Bachelet-Violette L, Rouzet F, Beilvert A, Autret G, Maire M, Menager C, Louedec L, Choqueux C, Saboural P, Haddad O, Chauvierre C, Chaubet F, Michel J-B, Serfaty J-M, Letourneur D (2014) Ultrasmall superparamagnetic iron oxide nanoparticles coated with fucoidan for molecular MRI of intraluminal thrombus. Nanomedicine 10(1):73–87. https​://doi.org/10.2217/ nnm.14.51 300. Xu J, Zhou J, Zhong Y, Zhang Y, Liu J, Chen Y, Deng L, Sheng D, Wang Z, Ran H, Guo D (2017) Phase transition nanoparticles as multimodality contrast agents for the detection of thrombi and for targeting thrombolysis: in vitro and in vivo experiments. ACS Appl Mater Interfaces 9(49):42525– 42535. https​://doi.org/10.1021/acsam​i.7b126​89 301. Tang T, Valenzuela A, Petit F, Chow S, Leung K, Gorin F, Louie AY, Dhenain M (2018) In vivo MRI of functionalized iron oxide nanoparticles for brain inflammation. Contrast Media Mol Imaging 2018:3476476–3476476. https​://doi.org/10.1155/2018/34764​76 302. Gu L, Li X, Jiang J, Guo G, Wu H, Wu M, Zhu H (2018) Stem cell tracking using effective selfassembled peptide-modified superparamagnetic nanoparticles. Nanoscale 10(34):15967–15979. https​://doi.org/10.1039/C7NR0​7618E​ 303. Naseroleslami M, Aboutaleb N, Parivar K (2018) The effects of superparamagnetic iron oxide nanoparticles-labeled mesenchymal stem cells in the presence of a magnetic field on attenuation

13

90

Reprinted from the journal

Topics in Current Chemistry (2020) 378:40

304. 305.

306. 307.

308.

of injury after heart failure. Drug Deliv Transl Res 8(5):1214–1225. https​://doi.org/10.1007/s1334​ 6-018-0567-8 Sweeney SK, Manzar GS, Zavazava N, Assouline JG (2018) Tracking embryonic hematopoietic stem cells to the bone marrow: nanoparticle options to evaluate transplantation efficiency. Stem Cell Res Ther 9(1):204–204. https​://doi.org/10.1186/s1328​7-018-0944-8 Garcés V, Rodríguez-Nogales A, González A, Gálvez N, Rodríguez-Cabezas ME, García-Martin ML, Gutiérrez L, Rondón D, Olivares M, Gálvez J, Dominguez-Vera JM (2018) Bacteria-carried iron oxide nanoparticles for treatment of anemia. Bioconjug Chem 29(5):1785–1791. https​://doi. org/10.1021/acs.bioco​njche​m.8b002​45 Khandhar AP, Wilson GJ, Kaul MG, Salamon J, Jung C, Krishnan KM (2018) Evaluating sizedependent relaxivity of PEGylated-USPIOs to develop gadolinium-free T1 contrast agents for vascular imaging. J Biomed Mater Res Part A 106(9):2440–2447. https​://doi.org/10.1002/jbm.a.36438​ Nimi N, Saraswathy A, Nazeer SS, Francis N, Shenoy SJ, Jayasree RS (2018) Multifunctional hybrid nanoconstruct of zerovalent iron and carbon dots for magnetic resonance angiography and optical imaging: an in  vivo study. Biomaterials 171:46–56. https​://doi.org/10.1016/j.bioma​teria​ ls.2018.04.012 Xiao Y, Lin ZT, Chen Y, Wang H, Deng YL, Le Elizabeth D, Bin J, Li M, Liao Y, Liu Y, Jiang G, Bin J (2015) High molecular weight chitosan derivative polymeric micelles encapsulating superparamagnetic iron oxide for tumor-targeted magnetic resonance imaging. Int J Nanomed 10:1155– 1172. https​://doi.org/10.2147/IJN.S7002​2

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Reprinted from the journal

91

13

Topics in Current Chemistry (2020) 378:12 https://doi.org/10.1007/s41061-019-0275-y REVIEW

Gold, Silver and Iron Oxide Nanoparticles: Synthesis and Bionanoconjugation Strategies Aimed at Electrochemical Applications Claudia Iriarte‑Mesa1 · Yeisy C. López1,2 · Yasser Matos‑Peralta1 · Karen de la Vega‑Hernández3 · Manuel Antuch4  Received: 31 July 2019 / Accepted: 13 December 2019 / Published online: 7 January 2020 © Springer Nature Switzerland AG 2020

Abstract Nanomaterials have revolutionized the sensing and biosensing fields, with the development of more sensitive and selective devices for multiple applications. Gold, silver and iron oxide nanoparticles have played a particularly major role in this development. In this review, we provide a general overview of the synthesis and characteristics of gold, silver and iron oxide nanoparticles, along with the main strategies for their surface functionalization with ligands and biomolecules. Finally, different architectures suitable for electrochemical applications are reviewed, as well as their main fabrication procedures. We conclude with some considerations from the authors’ perspective regarding the promising use of these materials and the challenges to be faced in the near future. Keywords  Gold nanoparticles · Silver nanoparticles · Iron oxide nanoparticles · Nanobioconjugation · Immobilization strategies · Biomolecules · Biosensors Abbreviations AFP Alpha-fetoprotein Chapter 4 was originally published as Iriarte‑Mesa, C.,·López, Y. C., Matos‑Peralta, Y., Vega‑Hernández, K. de la & Antuch, M. Topics in Current Chemistry (2020) 378: 12. https://doi.org/10.1007/s41061-0190275-y. * Manuel Antuch manuel.antuch‑cubillas@ensta‑paris.fr; [email protected] 1

Laboratorio de Química Bioinorgánica, Departamento de Química General e Inorgánica, Facultad de Química, Universidad de La Habana, Zapata y G, Vedado, Plaza de la Revolución, 10 400 La Habana, Cuba

2

Instituto Politécnico Nacional, Centro de Investigación en Ciencia Aplicada y Tecnología Avanzada, Calzada Legaria 694, Col. Irrigación, 11 500 Ciudad de México, Mexico

3

Sorbonne Université, CNRS, Institut Parisien de Chimie Moléculaire, 75005 Paris, France

4

Unité de Chimie et Procédés, École Nationale Supérieure de Techniques Avancées (ENSTA), Institut Polytechnique de Paris, 828 Boulevard des Maréchaux, 91120 Palaiseau, France





Reprinted from the journal

93

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:12

AgNPs Silver nanoparticles Apt Aptamer APTES (3-Aminopropyl)triethoxysilane AuNPs Gold nanoparticles CEA Carcinoembryonic antigen CFP-10 Culture filtrate protein CFU Colony forming unit CMX Carboxymethyl dextran CNT Carbon nanotubes CP Carbon paste CQDs Carbon quantum dots CS Chitosan CuAAC​ Cu(I)-catalyzed alkyne-azide cycloaddition DDT 1-Dodecanethiol DET Direct electron transfer DHEAS Dehydroepiandrosterone sulfate DMSO Dimethylsulphoxide DNA Deoxyribonucleic acid DPV Differential pulse voltammetry DTA/TG Differential thermal analysis/thermogravimetry EDC 1-Ethyl-3-(dimethylaminopropyl) carbodiimide EIS Electrochemical impedance spectroscopy EtOH Ethanol Fe3O4 NPs Magnetite nanoparticles GCE Glassy carbon electrode GC–MS Gas chromatography and mass spectrometry GO Graphene oxide GOx Glucose oxidase GPC-3 Glypican-3 GQD Graphene quantum dots Hc Coercitivity HCC Hepatocellular carcinoma HER2 Human Epidermal growth factor Receptor 2 HOBt  N-hydroxybenzotriazole HRP Horseradish peroxidase Ig Immunoglobulin IL Ionic liquid IONPs Iron oxide nanoparticles ITO Indium tin oxide LBL Layer by layer LOD Limit of detection microRNA Microribonucleic acid MNPs Magnetic nanoparticles Ms Saturation magnetization Mr Remanent magnetization MWCNT Multi-walled carbon nanotubes

13

94

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

NHS  N-hydroxysuccinimide NPs Nanoparticles OA Octanoic acid pACC​ Pediatric adrenocortical carcinoma PDVB Polydivinylbenzene PEDOT:PSS Poly(3,4-ethylenedioxythiophene) polystyrene sulfonate PEG Polyethylene glycol PEI Polyethyleneimine PPy Polypyrrole PSA Prostate-specific antigen PVA Polyvinylalcohol PVP Polyvinylpyrrolidone RGO Reduced graphene oxide RIL Rilpivirine RSD Relative standard deviation SAM Self-assembled monolayer SPR Surface plasmon resonance tB Brownian relaxation time tN Neel relaxation time TOAB Tetraoctylammonium bromide TOPO Tri-n-octyl phosphine oxide Tris Tris(hydroxymethyl)aminomethane WP Whatman filter paper ZnONRs Zinc oxide nanorods ΔG Variation of Gibbs free energy

1 Introduction Nanoscience and nanotechnology have allowed major breakthroughs in the development and use of new devices. Nanoparticles (NPs) have played an important role in the design of new materials for application in areas related to medicine [1, 2], water treatment [3, 4], electrochemical and optical sensing [5], catalysis [6–11] and electrocatalysis [12, 13]. Their outstanding position in current chemistry is due to the numerous advantages they confer, including biocompatibility and ease of functionalization [14], along with the possibility to make biosensors with increased sensitivity, reproducibility and reduced assay times [15, 16]. Among the many nanosystems now available, gold and silver NPs are the most well known, and possess a wide variety of interesting optical and electrochemical properties [17–20]. Likewise, iron oxide NPs have become widely employed because of their magnetic and electrochemical properties [21–24]. The excellent properties of NPs can be boosted by conjugation with biomolecules, allowing exploitation of the natural specificity of the biological component, while the NP scaffold provides increased stability and frequently improves the biological properties of the biomolecule as compared to its native state [25–29]. The wide variety of methodologies used to conjugate biomolecules to NPs guarantees innumerable applications of the bioconjugates obtained. Reprinted from the journal

95

13



Topics in Current Chemistry (2020) 378:12

Hence, selection of the most appropriate strategy constitutes a crucial aspect in the design and construction of functional nanobioconjugates [30, 31]. This contribution aims to highlight the virtues of conjugating diverse biomolecules with NPs. Herein, we focus on gold, silver and iron oxide NPs. Other nanometric systems such as quantum dots, carbon nanotubes, fullerenes, graphene, reduced graphene oxide (RGO), Janus and core–shell NPs are outside the scope of this review. The purpose of this contribution is to provide a pedagogic overview of the topic, with emphasis on the progressive presentation of the contents. Accordingly, we begin with simpler and more general concepts, such as the preparation, structural description and physicochemical characterization of the three kinds of NPs, later on covering the most advanced and up-to-date bioelectrochemical applications of NP–biomolecule conjugates. Our intention was to review the most recent advances in nanobioconjugation chemistry; consequently, 70% of the work reviewed here comes from the last 3  years (2017–2019), 86% from the last 5  years (2015–2019) and 93% from the last decade (2009–2019), thus assuring a current view of the most recent developments in the field. This review is organized into five sections in addition to this Introduction. The following section presents the general properties of gold, silver and iron oxide NPs. Thereafter, a section on  synthetic routes summarizes the main methods used to obtain these nanomaterials. We then discuss stabilization and surface functionalization of NPs with different functional groups, which allows their further bioconjugation to diverse biomolecules, as described in the next section, Conjugation of biomolecules to NPs: the main key to further application. Finally, Nanobioconjugates and their electrochemical application offers a review of the most recent trends in the design of electrochemical biosensors based on the conjugation of gold, silver, and iron oxide NPs with biomolecules. The use of different electrode architectures, and preparation protocols and biosensor performances are discussed, in order to provide a contemporary outline of current trends in the utilization of bioconjugated nanosystems in bioelectrochemical applications.

2 General Properties of Gold, Silver and Iron Oxide NPs 2.1 Gold and Silver NPs NPs are atomic aggregates with a diameter between 1 and 100 nm [32]. Such dimensions determine their unique properties, which differ from individual atoms and bulk materials. These differences are explained by their high surface to volume ratio and quantum size effects [33]. Both gold and silver NPs (herein referred to as AuNPs and AgNPs, respectively) present unique colors that have made them attractive for developing new analytical assays [34]. The colors of these colloids are caused by a phenomenon known as surface plasmon resonance (SPR) [35, 36]. This is a feature typical of many metallic NPs, including AuNPs and AgNPs. SPR consists of the collective oscillation of conduction electrons across the NP due to the resonance that occurs with incident radiation (Fig. 1).

13

96

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

Fig. 1  Schematic representation of surface plasmon resonance (SPR) in metallic nanoparticles (NPs). The sinusoidal line represents the visible electromagnetic wave

In 1908, Gustav Mie provided the explanation for the visible-light absorption of colloidal gold suspensions. Mie assumed small spherical and homogenous NPs interacting with an electromagnetic field in order to solve Maxwell’s equations [37]. For NPs with a diameter comparable to the excitation wavelength, the extinction coefficient Cext is governed by the dipolar absorption term in Mie’s equation. This case is known as the quasi-static or dipolar approximation (Eq. 1) [38].

Cext

3∕ 𝜀i 24𝜋 2 R3 𝜀m 2 = ( )2 𝜆 𝜀r + 2𝜀m + 𝜀2i

(1)

where R is the radius of the spherical NP, 𝜆 is the wavelength, 𝜀m is the dielectric constant of the medium related to the refraction index through the expression 𝜀m = 𝜂m2 . The terms 𝜀r (𝜔) and 𝜀i (𝜔) represent the real and imaginary component of the dielectric function of the nanoparticle (Eq. 2). [ ] 𝜀(𝜔) = 𝜀r (𝜔) + i𝜀i (𝜔) . (2) Resonance is produced when 𝜀i (𝜔) is small, or depends poorly on the oscillation frequency 𝜔 and, therefore, Eq. 3 holds.

𝜀r (𝜔) = −2𝜀m

(3)

The previous analysis describes the SPR maximum behavior only for NPs below 10 nm. However, the size of the NPs influences the position of the SPR due to the dependence of the dielectric constant of the metal on NP size. The dielectric constant of the material may be adjusted for different sizes by introducing a relaxation frequency parameter. Amendola et  al. [39] developed a procedure to fit UV–Vis spectra for AuNPs using Gans’ model (an extension of Mie’s theory for nonspherical particles [40]) considering only the average radius of the NPs (R), a standard deviation with a Gaussian distribution, and the fraction of spherical to spheroidal AuNPs. The method allows the size, concentration, and aggregation level of AuNPs to be estimated with an accuracy of 6% for AuNPs between 4 and 25 nm. This procedure Reprinted from the journal

97

13



Topics in Current Chemistry (2020) 378:12

may be extended to other nanometric systems such as AgNPs and noble metal alloys [40]. For AuNPs and AgNPs, the resonance condition is satisfied at visible wavelengths. Thus, AuNPs of around 10 nm are red, possessing a maximum at approximately 510  nm; whereas AgNPs of similar size are yellow, corresponding to an absorption maximum at around 400  nm. Multiple factors influence the maximum of the SPR, including metal type [41], NP size [42], shape [43], functionality of the nanostructure [44], composition [45], interparticle distance [46], ligand–NP interactions [47], the refractive index of the medium [48], pressure [49] and temperature [50]. 2.2 Iron Oxide NPs Among metal oxides NPs, iron oxide NPs (IONPs) have attracted special interest due to their attractive magnetic properties as well as their high biocompatibility and low toxicity [51]. Magnetic materials are classified as ferromagnetic, ferrimagnetic, antiferromagnetic and superparamagnetic (Fig. 2). Ferromagnetic materials present multiple magnetic domains, where the electron spins are oriented in the same direction and therefore display a remanent macroscopic magnetization. In ferrimagnetic materials, a portion of the magnetic moments inside the domains are oriented in one direction and the rest in the opposite direction. Thus, in each domain, there is a magnetic dipole. These materials show remanent magnetization after the removal of an external magnetic field, but smaller than that observed for ferromagnetic materials. When the number of magnetic moments oriented in opposite directions is equal, the net magnetic moment is null, and, as a consequence, the material is antiferromagnetic. Finally, in a superparamagnetic material, the magnetic dipoles align themselves in the direction of an external magnetic field. By withdrawing the external

Fig. 2  a Representation of the domain structure of ferromagnetic, antiferromagnetic and ferromagnetic solids. b Spin orientation inside the magnetic domain at Ms and Hc = 0 for a superparamagnetic solid. c Coercivity (Hc) as particle diameter (D) function

13

98

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

field, thermal fluctuations cause disordered movement of the magnetic moments, which causes a null magnetic moment. The superparamagnetic behavior of a material is the result of magnetic anisotropy. Two main contributions to anisotropy exist, related to (1) crystal anisotropy and (2) particle shape and morphology. Crystal anisotropy is associated with the existence of different crystallographic directions along which the spins of the electrons are aligned faster, and, therefore, the substance is easily magnetized. The other anisotropic contribution is related to the size and morphology of the particles; for example, spherical particles are prone to present smaller coercivity than elongated particles [52]. If the material size decreases toward the nanoscale, the number of spins that can be oriented by an external magnetic field is increased. In this case, a tendency toward superparamagnetic behavior is produced. It is known that IONPs can adopt a superparamagnetic behavior above a critical diameter (Fig. 2c), i.e., the NPs are attracted to the magnet but when the magnetic field is removed they lose magnetization [53]. The critical diameter where superparamagnetic behavior is observed is around 15–20 nm for magnetite NPs [54]. Key magnetic parameters used to assess the potential application of those systems are (1) saturation magnetization (Ms), (2) remanent magnetization (Mr), (3) the coercivity (Hc) and (4) the Neel and Brownian relaxation time of nanoparticles (­tN and ­tB, respectively) as illustrated in Fig. 3. The saturation magnetization (Ms) can be defined as the highest value of magnetization reached by the material. This value corresponds to the parallel alignment of all magnetic moments of the system. When the magnetic field is diminished, the magnetization (M) at H = 0 is not equal to zero, and this is called remanent magnetization (Mr). On the other hand, the value of the magnetic field at which the magnetization becomes zero is known as coercivity (Hc). In the case of nanoparticles, superparamagnetic behavior is characterized by nearly zero Hc and Mr. Another distinguishing feature of magnetic NPs (MNPs) is their capacity to generate heat when the NPs are exposed to a high-frequency magnetic field. Under these conditions, the monodomain nanoparticles generate heat through the oscillation of their magnetic moment. This heat may be dissipated via two different processes, (1) Brownian relaxation, produced by the rotation of the entire magnetic particle within a surrounding liquid, and (2) Néel relaxation, which is the rotation of the magnetic

Fig. 3  a Magnetization curve (M vs. H) for a magnetic solid, showing Ms, Mr and Hc. b Magnetization curve for superparamagnetic nanoparticles showing nearly zero Mr and Hc Reprinted from the journal

99

13



Topics in Current Chemistry (2020) 378:12

moment within the magnetic core. In most cases, the heating power is governed by the fastest regime of relaxation [52]. The ability of IONPs to generate heat in the presence of an alternating magnetic field [55] has made them the base for the development of hyperthermia applications, employed mainly to induced death in tumor cells [56]. In order to produce hyperthermia, energy dissipation phenomena must be involved, according to Néel or Brown relaxation [57], although in most cases is difficult to distinguish which process is actually taking place. The most studied iron oxides at the nanoscale are hematite (α-Fe2O3), magnetite ­(Fe3O4) and maghemite (γ-Fe2O3). Hematite presents a hexagonal unit cell with a = 0.5034 nm and c = 1.375 nm, and six formula units per cell. Hematite may also appear in the rhombohedral system with a = 0.5427 nm and α = 55.3º, and two formula units per cell. The structure consists in an hcp array of oxygen ions along the [001] direction and two-thirds of the sites are occupied by Fe(III) arranged regularly, with two filled sites followed by one vacant site in the (001) plane. At bulk scale, hematite behaves as a weak ferromagnetic material or antiferromagnetic. On the other hand, magnetite and maghemite present a cubic unit cell, and, and at bulk scale behave as a ferrimagnet. The structure of magnetite is typical of an inverse spinel containing iron(II) and iron(III) distributed in its structure. Fe(III) ions are located in both tetrahedral and octahedral sites, while Fe(II) ions occupy only octahedral sites due to the higher ferrous crystal field stabilization energy. The cubic unit cell has an edge length a = 0.839 nm, presenting eight formula units per cell. Magnetite is non-stoichiometric, with cation deficiency in the Fe(III) sublattice. Maghemite is isostructural to ­Fe3O4, but the majority of Fe ions are trivalent, and cation vacancies compensate the oxidation of Fe(II). Maghemite has a cubic unit cell with a = 0.834 nm. In the cell, eight cations occupy tetrahedral sites and the rest are distributed randomly in octahedral sites. The vacancies are limited to octahedral sites [58].

3 Synthetic Routes 3.1 Synthesis of AuNPs There are numerous reports on the different methodologies for the synthesis of AuNPs with a wide variety of sizes and morphologies (reviewed in [59, 60]). Herein, we will describe only the most important. The chemical synthesis of AuNPs is based on the reduction of a gold salt and the stabilization of the NPs. The two conventional chemical methods for obtaining AuNPs are the Turkevich and the Brust-Schiffrin methods. In 1951, Turkevich reported an easy pathway to obtain spherical AuNPs by reducing hydrogen tetrachloroaurate(III) ­(HAuCl4) using trisodium citrate as both reducing agent and stabilizer [61]. This method involves the use of water as a solvent, and the citrate is added once the solution is boiling. The NPs obtained are in a range from 10 to 150 mn in size. The Turkevich method is still used widely due to its simplicity and reproducibility. Recent studies show that citrate can be further substituted for α and

13

100

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



β-hydroxy carboxylates as stabilizers in order to prepare nanoparticles of various sizes, size distribution, and shape [62]. The Brust-Schiffrin approach [63] was reported in 1994 and presented a new procedure for the synthesis of spherical NPs somewhat similar to the Turkevich method, but the size of the particles obtained was below 10  nm. In this method, metallic clusters are grown at the same time as a self-assembled thiol monolayer is attached to the growing nuclei. In order to achieve both processes simultaneously, the reaction occurs in two phases (water and toluene). Sodium borohydride (­ NaBH4) is used as a reducing agent and tetraoctylammonium bromide (TOAB) as a phase-transfer catalyst. The gold complex anion, [­ AuCl4]−, is transferred from the aqueous phase to the organic phase by electrostatic interactions between the positively charged TOAB and the negative gold ion. Seed-mediated growth method is another strategy to obtain AuNPs. Unlike the above-mentioned methodologies, seed-mediated growth is one of the easiest ways to effectively control the size of the NPs [64, 65]. It was first reported by Wiesner in 1989 [66] and consists of obtaining gold seeds via reduction of ­HAuCl4, and then adding more H ­ AuCl4, which is reduced using phosphorus. This methodology has evolved, and is currently used for the preparation of AuNPs with narrow size distribution [67]. Other strategies report the use of fungi [68] and plant extracts [69] as reducing agents; photosynthesis [70], and laser ablation [71] are additional approaches for preparing colloidal gold, as illustrated in Fig. 4. 3.2 Synthesis of AgNPs AgNPs can be synthesized by reducing silver ions in aqueous or non-aqueous solutions through chemical methods. For example, chemical reduction with sodium citrate [72], sodium borohydride [73], and calcium ascorbate [74], have been used as

Fig. 4  Most common synthetic routes for preparing gold NPs (AuNPs) and silver NPs (AgNPs) Reprinted from the journal

101

13



Topics in Current Chemistry (2020) 378:12

reducing and stabilizing agents. Green synthesis of AgNPs has become very trendy in the past few years [75]. This methodology has gained increasing attention due to its eco-friendliness and low cost [76]. There is a great variety of plants that have been reported as agents for AgNPs synthesis under mild conditions [75, 77]. Some of the most common plant extracts that can behave as reducing agents and stabilizers are from Polyalthia longifolia [78] or Moringa oleifera [79]. Some families of bacteria, for example Pseudomonas [80], have been used for the preparation of colloidal silver. Other methodologies for obtaining AgNPs (common to AuNPs) are based on microemulsion techniques [81], microwave heating [82], and laser ablation [83] (Fig. 4). 3.3 Synthesis of IONPs As described for AuNPs and AgNPs, IONPs can be synthesized through a large diversity of methods, and these can be categorized into physical, chemical and green syntheses. Chemical methods are most often used, particularly chemical co-precipitation and thermal decomposition. In 1981, Massart first reported the preparation of magnetite NPs in alkaline media using ­FeCl3 and ­FeCl2 as precursors in a molar ratio 1:2 ­(Fe2+/Fe3+) under inert atmosphere (Eq. 4) [84].

Fe2+ (ac) + 2Fe3+ (ac) + 8OH− (ac) → Fe3 O4 (s) + 4H2 O.

(4)

This procedure allows a large mass of NPs to be obtained but with broad size distribution and low crystallinity. Unlike chemical coprecipitation, thermal decomposition of organic complexes produces NPs with narrow size distributions and high crystallinity, which makes it the most attractive method to prepare IONPs for biomedical applications [85, 86]. This method is based on the thermal decomposition of organic complexes of iron, such as iron oleate [87], pentacarbonyliron(0) [88], Fe(acac)3 [89] or ferrocene [90]. Other chemical strategies are sol–gel [91], oxidation [92], and solvothermal reactions [93]. The laser ablation technique allows preparation of NPs with various sizes and compositions by modifying the wavelength, pulse duration, and power of the laser [94]. On the other hand, green synthesis of IONPs has gained much attention recently due to their good reproducibility, low cost, high yields, mild reactions conditions and biocompatibility [95].

4 Stabilization and Surface Functionalization 4.1 Functional Groups Relevant for NP Modification In the case of noble metal NPs (e.g., Au and Ag), thiolated ligands are essentially used to cover the surface due to the strong Au–S and Ag–S interaction (approximately 200 kJ mol−1 [96]). In the case of IONPs, the most employed coating agents are carboxylates, phosphates, hydroxyls, and in some cases thiols, although oxygenated groups are preferred for modifying IONPs [97] (Fig. 5). The choice of the

13

102

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



Fig. 5a–k  Functional groups with high affinity to the surfaces of NPs. a Thioesters, b thiols, c dithiocarbamates, d thioureas, e phosphine oxides, f amines, g phosphates, h catechins, i trimethoxysilane, j carboxylic acids, k alcohols

strategy to stabilize and functionalize NPs depends strongly on the envisaged application. Certainly, if the final application of the NPs is a biosensor, the preferred ligand will be a biomolecule capable of specifically recognizing a molecule of interest. If the aim is to provide biocompatibility, ligands with biological significance are preferred, such as polyethylenimine (PEI), carboxymethyl dextran (CMX) or polyethylene glycol (PEG) [98]. 4.2 Ligand Exchange Ligand exchange consists of replacing the coating agent of the NP with another coating agent presenting higher affinity toward the NP surface. This process may be performed right after the synthesis of the NPs. Kluenker and co-workers conducted a study for the thiol exchange process on AuNPs coated with oleylamine [99]. The ligand exchange process was followed by solution NMR. When oleylamine was replaced by 1-octadecanethiol, the tests showed that the procedure corresponded to an equilibrium reaction. This study reveals how complex the ligand exchange process is, and that efficient surface coverage depends on the repeated exchange reactions with large ligand excess, the size of NPs, and the size of the ligand. The authors found that a more effective ligand replacement is achieved after repeated functionalization reactions using high ligand concentration or by reducing NP size [99]. Moreover, the cellular uptake of AuNPs through ligand exchange has been reported recently; 12 different ligands (small thiolated molecules, thiolated Reprinted from the journal

103

13

Topics in Current Chemistry (2020) 378:12

polymers, thiolated and non thiolated biomolecules) were tested, showing that cell uptake is influenced by the effect of the exchangeable ligand and thus the NPs can aggregate within the lipid bilayers. These results represent a step forward in our understanding of the cellular uptake and cell integrity once NPs are incorporated into the cell [100]. For biological applications, it is crucial to know the composition of the surface precisely. Locardi et  al. [101] prepared 1-dodecanethiol-coated AuNPs for a posterior ligand exchange using 11-mercaptoundecanoic acid. The latter authors achieved partial substitution of the hydrophobic coating for a more hydrophilic one, thus improving the water stability of the NPs for further biological applications. The layer composition of the NPs obtained by these researchers was determined by differential thermal analysis/thermogravimetry (DTA/TGA) coupled with gas chromatography and mass spectrometry (GC–MS) [101]. AgNPs coated with polyvinylpyrrolidone (PVP) were prepared by Sang Cho and co-workers [102]. Afterward, the capping agent was replaced using propanethiol. These authors obtained NPs with very low aggregation tendency and high thermal stability. The final purpose was to incorporate more stable AgNPs inside perovskites for solar cell applications [102]. The effect of ligand exchange on the crystallite size of AgNPs coated with oleic acid was further evaluated by Okada [103]. In this latter work, three ligands were tested: tri-n-octyl phosphine oxide (TOPO), octanoic acid (OA), and 1-dodecanethiol (DDT). The highest increase in crystallite size of AgNPs was observed for TOPO as the exchangeable ligand [103]. Likewise, to modify the surface, IONP ligand exchange may also be carried out. A new and precise synthetic route to obtain magnetite NPs (­Fe3O4 NPs) functionalized with polymeric ligands through ligand exchange has been reported recently [104]; this produced NPs with different surface charges, allowing evaluating their interaction with cells.

5 Conjugation of Biomolecules to NPs: the Main Key to Further Applications The application of NPs requires the stabilization of colloidal systems in a polar medium, especially water, and the subsequent conjugation of biomolecules. Several biomolecules, such as proteins [105], polypeptides [106] and antibodies [107] have been conjugated to the surface of nanomaterials. Conjugation allows to combine the biological functionality of the biomolecule with the chemical and mechanical stability of the support, it guarantees the increase in operational stability, facilitates the separation and purification of biomolecules and increases their reusability [108, 109]. However, conjugation also causes alteration of the conformation of the biomolecule, which may cause the loss of biological activity and the presence of fractions of immobilized macromolecules with a different number of junctions to the support [110, 111]. In the case of immobilized enzymes, despite the fact that conjugation methodologies are expensive, the possibility of recovering biocatalysts from the reaction medium and the development of continuous operations reduce the overall costs of the process [112].

13

104

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

The crucial parameters to consider during the design of functional bioconjugates are (1) the dispersion of the colloidal system in an aqueous medium, (2) the stability of the association of the biomolecule with the support, and (3) the orientation and functionality of the conjugated biomolecule [113, 114]. The most commonly used strategies for bioconjugation include physical adsorption [115], covalent binding [116], inclusion inside polymeric matrices [117] and conjugation through supramolecular interactions based on molecular recognition [118] (Fig. 6). 5.1 Physical Adsorption Physical adsorption occurs through the interaction between the biomolecule and the nanometric support via multivalent interactions, including van der Waals forces, the formation of hydrogen bonds, hydrophobic interactions and electrostatic interactions, depending on the traits of the biomolecule and the coating of the NPs [119]. Bioconjugation through physical interactions is highly versatile, since it offers the possibility to separate the biomolecule from the NPs by changing the conditions of the medium, which has allowed the use of these methods in the design of composites for drug delivery and controlled release of drugs [120]. However, this strategy lacks control of the spatial orientation of the immobilized biomolecules, which may lead to the decrease or loss of their biological activity [121]. Due to the reversible

Fig. 6  Strategies of bionanoconjugation Reprinted from the journal

105

13



Topics in Current Chemistry (2020) 378:12

character of the physical interactions, variations in pH [122], temperature [123] or polarity of the medium [124] may promote the desorption of the immobilized component. For electrostatic adsorption, the conjugation of biomolecules to NPs relies on the attraction due to opposite charges in both species. The surface of NPs can be engineered to have a specific charge in order to promote the interaction with biomolecules. For example, the reversible immobilization of trypsin has been reported through the direct immobilization of the enzyme via the electrostatic interactions between the positively charged trypsin and the negatively charged citric acid coating ­Fe3O4 NPs [125]. The adsorption of proteins onto inorganic supports, such as noble metal and metal oxide NPs, and the influence of electrostatic interactions in the conjugation, have been described in several reports [126–128]. These interactions have proved to be particularly useful in the assembly of plasmid deoxyribonucleic acid (DNA) onto the surface of MNPs. Bioconjugation is promoted by the strong interaction between the negative charge associated with the phosphate backbone of most nucleic acids and the positively charged F ­ e3O4 NPs coated with a layer of silicon dioxide [129]. This effect may be achieved with other ligands possessing amino groups, such as (3-aminopropyl)triethoxysilane (APTES), chitosan and tris(hydroxymethyl)aminomethane (Tris) [130]. Despite the simplicity and efficiency of this strategy of conjugation, it is highly sensitive to the pH and the ionic strength of the medium [131]. On the other hand, hydrophobic interactions are commonly used to adsorb hydrophobic drugs onto nanometric supports. In this case, the surface of the NPs is modified with hydrophobic molecules to allow the absorption of the drugs, which are then triggered and released inside cells when the coatings of the NPs are degraded. Wu et  al. [132] described the synthesis of Janus nanoparticles comprised of Au nanorods and a polydivinylbenzene (PDVB) matrix, as a promising biomedical material for cancer treatment. The hydrophobic components of PDVB were used as carriers of curcumin for chemotherapy. This anticancer drug was loaded into the Janus nanoparticles due to strong hydrophobic interactions between curcumin and PDVB. The UV–Vis extinction of Janus NPs in the near infrared region also makes them an ideal candidate for photothermal therapy, guaranteeing significant decrease of cell viability, migration, and invasion as a result of combined chemo- and photothermal-effects [132]. 5.2 Supramolecular Conjugation Supramolecular conjugation is based on the molecular recognition between a ligand and a receptor, in which both the ligand and the receptor accommodate each other in a dynamic process based on chemical complementarity, without the actual formation of covalent bonds. When NPs are involved, the receptor is generally located at the surface of the NP, and immobilization occurs by means of a reaction via host–guest self-assembly [133]. Supramolecular interactions are also reversible, but possess the advantage of directing the orientation of biomolecules, which is the main advantage of the supramolecular strategy as compared to

13

106

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

covalent immobilization or physical adsorption. Several reports demonstrate that the conformation of enzymes is not heavily harmed when supramolecular methodologies are used [134, 135]. Due to its importance, we highlight in this revision the supramolecular immobilization using the biotin-avidin interactions. Certainly, avidin is a glycoprotein having four identical subunits that have high specificity and affinity for biotin, resulting in a strong interaction (Kaff ~ 1015 [136]). Streptavidin, a non-glycosylated tetramer protein, is the most commonly used avidin analog for immobilization through supramolecular interactions [137]. Biotinylated proteins [138], oligonucleotides [139], antibodies [140], and single-stranded DNA [141] have been immobilized on AuNPs and nanohybrids functionalized with streptavidin to obtain biosensors (Fig. 7). This methodology has also been used to conjugate gold/iron oxide core–shell nanoparticles with both chloroperoxidase and glucose oxidase (GOx) through layer-by-layer assembly using the specific avidin–biotin interactions. The bienzymatic nanoreactor obtained combined the generation of H ­ 2O2 in situ from glucose and oxygen, with facile separation using a magnetic field. The latter had a positive impact on the reusability of the obtained system. The combination of both enzymes also enhanced considerably the operational stability of the bioconjugate [142]. Other molecules, such as the family of cyclodextrins, also behave as hosts, and therefore allow the supramolecular conjugation of NPs caped with cyclodextrins with adamantane-functionalized biomolecules via inclusion complexes [143]. Other molecules such as ferrocene have also been used as cyclodextrin guests for the functionalization of modified biomolecules [144]. Assembly of a thiolated cyclodextrin and a ferrocene derivative has also been reported. The NPs obtained were used for controlled drug delivery. The noncovalent complex formed between cyclodextrin and an amphiphilic ferrocene derivative allowed encapsulation of doxorubicin; the drug may be rapidly released at low pH [145].

Fig. 7  Schematic representation of the supramolecular interaction between biotinylated biomolecules with NPs capped with streptavidin or avidin Reprinted from the journal

107

13



Topics in Current Chemistry (2020) 378:12

In addition, enzymes such as laccases have been immobilized in AuNPs by supramolecular interactions between the aromatic amino acids of the side chains of the enzyme and functionalized ferrocene molecules as a coating on the NPs. The results were compared with an analogous covalent coupling, demonstrating the advantages of supramolecular conjugations [146, 147]. Antibody–antigen interactions have also been used for bionanoconjugaion. In several cases, immobilization is beneficial to increase the nanoparticle-antibodyantigen association constant, relative to the free antibody [148, 149]. The immobilization of antibodies on nanomaterials has found several applications in diagnosis [150], biosensing [151], magnetic separation [152], purification of analytes [153] and cell labeling [154]. 5.3 Encapsulation Encapsulation on nanometric supports occurs when NPs coated with hydrophobic ligands are overcoated with amphiphilic ligands. The process at the surface of the NPs is achieved by intercalating the hydrophobic portions of ligands (or biomolecules) of the solution with the hydrophobic coatings of the NPs. Thus, the hydrophilic portion of the immobilized component (ligand or biomolecule) is oriented towards the solution. This methodology allows the stabilization of colloidal systems in hydrophilic media due to the head groups within the hydrophilic portion of the coating. In addition, the functional group may also allow further bioconjugation of the NPs, as previously reported [155]. This methodology allows the immobilization of drugs inside polymeric nanoparticles to develop controlled release agents and nanodelivery systems in cancer therapy [156]. Encapsulation of essential oils in poly(ɛ-caprolactone) nanocapsules has been reported to enhance the antimicrobial activity against food-borne pathogens. This is promising for food preservation [157]. Essential oils have also been encapsulated using several supports with different designs including NPs [158], nanocapsules [159], nanoemulsions [160], micelles [161], and liposomes [162]. In these cases, the formation of polymer-based and lipid-based nanosystems avoids the drawbacks of essential oils related to their volatility and low solubility, and also enhances antioxidant, anti-inflammatory and antibacterial activities of the encapsulated compounds. Nanoencapsulation has been also used to enable the production of “healthy” foods and drugs, through delivering specific substances such as vitamins [163], antibiotics [164], flavors [165], antioxidants [166], omega-3 fatty acids [167], proteins [168], and nucleic acids [169] in an easily absorbable nanometric form. The operational lifetime of enzymes for environmental applications may be enhanced through encapsulation. As an example, laccase encapsulated within chitosan NPs showed temperature and pH activity profiles similar to those of free enzyme, but the procedure of encapsulation ensured its stability against microbial degradation, allowing clear applications in industrial bioremediation [170]. Encapsulation has also extended the use of enzymes in conditions where these biocatalysts are neither stable nor active. For instance, GOx has been encapsulated within polymeric nanocapsules with a hydrophilic core via inverse miniemulsion periphery

13

108

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



RAFT polymerization. The encapsulated enzyme showed high enzymatic activity and stability in a mixture of toluene/t-BuOH, in which the free enzyme undergoes denaturation [171]. 5.4 Chemical Conjugation Chemical bioconjugation allows the immobilization of biomolecules at the surfaces of nanoparticles by forming covalent bonds. This method guarantees stable anchoring between the support and the biomolecule. Several works report that supramolecular conjugation, encapsulation or adsorption, are simpler and more economical methods to immobilize enzymes onto NPs, but show long-term loss of activity since the enzyme-support binding is weak compared to covalent methods [172, 173]. Covalent immobilization ensures that the number of immobilized molecules remains constant after conjugation with a small probability of losing the biomolecule during operation of the nanobioconjugate. For conjugates obtained by covalent immobilization of enzymes, greater resistance to deactivation due to the effects of temperature, organic solvents or pH has been reported [174, 175]. Although the formation of covalent bonds can affect the active conformation of enzymes, several reports describe the covalent coupling of these biocatalysts onto metal oxide NPs as a method for ensuring the increase in the enzymatic activity and stability of the immobilized biomolecule [176, 177]. The coupling protocol is based on reactions involving amino (–NH2) [178], carboxyl (–COOH) [179], hydroxyl (–OH) [180], azide ­(N3−) [181] or thiols (–SH) [182]—functional groups present on the surface of the support as well as in the side chains of the biomolecule. Of the 20 amino acids, the ones used most often for the formation of covalent bonds with the support are those that present ionizable groups in their side chain of biomolecules, such as lysine [183], cysteine [184], tyrosine [185], histidine [186], arginine [187], and aspartic [188] and glutamic acids [189]. One of the most widely used protocols for the chemical immobilization is known as the Steglich reaction, or the carbodiimide method, which is based on peptide coupling between a carboxyl group and a primary amine [107, 190]. During the chemical reaction, the carboxylic acid is activated in the presence of a coupling agent such as 1-ethyl-3-(dimethylaminopropyl) carbodiimide (EDC), forming an activated ester. Frequently, additives such as N-hydroxybenzotriazole (HOBt) or N-hydroxysuccinimide (NHS) are added to increase yields and decrease side reactions. Finally, the amine reacts with the activated ester to form an amide, with excellent yields (> 90%) (Scheme 1) [191]. The described methodology should be optimized carefully to avoid any irreversible damage to the biological function of the immobilized component. If the covalent immobilization is properly designed, the spatial disposition of the biomolecule can be controlled [192]. Alternatively, it is possible to use the high reactivity and selectivity of the isothiocyanate group towards primary amines to form thioureas [193]. This protocol has been used in biochemistry for the conjugation of antibodies [194], as well as in the labeling of nanoparticles coated with amino groups using fluorescent isothiocyanates for

Reprinted from the journal

109

13

Topics in Current Chemistry (2020) 378:12

Scheme 1  Formation of an amide by covalent coupling with the 1-ethyl-3-(dimethylaminopropyl) carbodiimide (EDC) and N-hydroxysuccinimide (NHS) (Steglich) reaction

biomedical applications [195]. The reaction is practically quantitative, and it is possible to use dimethylsulphoxide (DMSO) or ethanol (EtOH) as solvents (Scheme 2) [196]. Another route commonly used in bionanoconjugation is the alkylated azide cycloaddition catalyzed by Cu (I), which is based on the concept of “click chemistry”. Cu(I)catalyzed azide/alkyne 1,3-dipolar cycloadditions, also known as “CuAAC’’, involve the cycloaddition of an alkyne and an azide to form a 1,2,3-triazole ring, which is a strong linker between the NP and the biofunctional agent (Scheme 3) [197]. Although bonds formed by click chemistry are electronically similar to amide bonds, triazols are highly stable whereas amide bonds, and also disulfide linkages formed by other direct conjugation strategies, are prone to cleavage by hydrolysis and reduction, respectively [198]. The reaction has demonstrated high versatility and suitability for the conjugation of several chemical compounds, from small molecules to proteins [199]. This methodology is characterized by specific conjugation, as azide and alkyne reactive groups are highly specific to each other and do not react with most other functional groups. “CuAAC’’ reactions are commonly used despite the fact that neither alkynes nor azides are present as functional groups in naturally occurring biomolecules. This guarantees highly oriented linkages, and control of the disposition of the immobilized biomolecule [200]. Click chemistry has been implemented with gold nanoparticles functionalized with a synthetic functional copolymer consisting of a backbone of polydimethylacrylamide (DMA) functionalized with an alkyne monomer. The polymeric coating guaranteed the stabilization and functionalization of the colloidal system, as well as the alkyne functionalities for the interaction with azido groups from modified anti-mouse IgG antibody. NPs functionalized with antibodies were applied to the development of gold labels for biosensing applications [201].

Scheme 2  Formation of thioureas by the reaction of the isothiocyanate group with a primary amine

13

110

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



Scheme 3  Mechanism of formation a 1,2,3-triazole by a click chemistry reaction

Overall, it becomes clear that the conjugation of NPs with macromolecules is extensive. Several possibilities have been explored, including polymers, surfactants, organic ligands, biological ligands, amino acids, peptides, DNA strains, antibodies, aptamers, and proteins, with multiple inorganic NPs, aiming for a wide range of applications as depicted in Fig.  8. Among these multiple applications, we have selected those related to bioelectrochemistry to illustrate the use of the nanobioconjugation strategies described so far in practical cases.

6 Nanobioconjugates and their Electrochemical Applications In general, metal NPs have excellent conductivity and catalytic properties, which make them suitable for acting as “electronic wires” to enhance electron transfer between redox centers in proteins and electrode surfaces [202]. Electrodes modified with NPs show electrochemical responses that depend on the size of the NPs, the space between the particles and the particle density [203, 204]. Furthermore, the molecular components of NPs have a strong influence Reprinted from the journal

111

13



Topics in Current Chemistry (2020) 378:12

Fig. 8  Surface functionalization of NPs with different biomolecules

on the electron transfer rate as a consequence of interfacial changes resulting in electron tunneling through NP–NP and NP–electrode junctions [204]. It is convenient to highlight that the electrochemical interpretation of bionanoconjugated systems is delicate, and results may be misinterpreted. Bartlett et al. [205] argue that the vast majority of articles reporting biosensors with direct electron transfer (DET) between native GOx and nanostructured electrodes are incorrect. This is due to the lack of a critical and deep analysis of experimental evidence. Bartlett and coauthors demonstrate their hypotheses by an exhaustive study of experimental results, and suggest evidence that should be fulfilled in order to support any serious claim of DET for GOx. Several architectures are reported in the literature to conjugate biomolecules with nanoparticles, such as a totally random distribution [206], a self-assembled monolayer (SAM) [147] or a layer-by-layer (LBL) assembly [174] (Fig. 9). The main methods summarized in the section  Conjugation of Biomolecules to NPs: the Main Key to Further Applications are also employed for the preparation of electrochemical biosensors, viz. adsorption [207], encapsulation [208], covalent binding [209], cross-linking [210] and supramolecular associations [211]. Herein, we will focus on major issues regarding bioconjugates of AuNPs, AgNPs and IONPs for bioelectrochemical applications.

13

112

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



Fig. 9  Possible architectures obtained when the surface of the electrode is modified with biomolecules

6.1 Electrochemical Applications of Bioconjugates Based on AuNPs AuNPs are attractive scaffolds due to their compatibility with the immobilization of biomolecules while maintaining catalytic activity and providing a large specific surface area and good electrocatalytic properties [212, 213]. For instance, the use of AuNPs for the development of glucose biosensors has been exploited extensively due to good electrical communication between the transducer and the active redox site of the enzyme [214]. Rassas et al. [212] recently reported the development of a nanobiosensor for glucose detection capable of detecting concentrations of the order 5 µM and a dynamic range between 10 µM and 7 μM. Here, the presence of AuNPs improved the performance of biosensor based on a mixed polysaccharide/GOx composite. This biosensor established improvements such as avoiding loss of selectivity, high sensitivity (283.9 µA/log[glucose]), allowing the application to real biological samples such as saliva and serum [212]. Zou et al. [215] developed an electrochemical immunosensor for enzyme-free detection of E.coli K12 by employing as nanocomposite AuNPs attached to the polypyrrole (PPy)-RGO) surface through electrostatic adsorption. This serves as

Reprinted from the journal

113

13

Topics in Current Chemistry (2020) 378:12

a platform for the immobilization of a capture antibody, which was conjugated onto a composite made of ferrocene and doped polypyrrole-AuNPs. The fabricated immunosensor showed a linear range from 1.0 × 101 to 1.0 × 107  colony forming unit (CFU) ­mL−1 and a low detection limit of 10 CFU mL−1. Hoo et  al. [216] recently reported the construction of an electrochemical glucose biosensor based on a nanocomposite of zinc oxide nanorods (ZnONRs) and AuNps. In another example, AuNPs were then drop-casted on ZnONR/ITO substrates, while the GOx enzyme was immobilized with a matrix of Nafion onto the modified electrode. The results suggest direct electron transfer from GOx enzymes to the modified electrode. Moreover, the use of AuNPs can provide an additional pathway that facilitates electron transfer. The optimized conditions for this bioelectrode were Nafion/60 µL GOx/80 µL 30 nm AuNP/ZnONR/ITO, exhibiting high sensitivity of 14.53 and 2.54 µA ­mM−1 ­cm−2 for a wide working range of 0.05–1.0 and 1.0–20 mM and a low limit of detection (LOD) of 0.18 mM [216]. Buk and co-workers [217] reported a nanohybrid system based on carbon quantum dots (CQDs) and AuNPs for the design of enzymatic electrochemical biosensors using the CQDs/AuNPs as an immobilization matrix for a large surface area microfabricated gold electrode. As a proof of concept, the GOx enzyme was chosen as a model system, and immobilized onto the microfabricated gold electrode surface. GOx was conjugated a CQDs/AuNPs nanohybrid by cross-linking with glutaraldehyde. The biosensor exhibited high electrocatalytic activity, a good sensitivity of 47.24 µA ­mM−1 ­cm−2, reproducibility (5.4% relative standard deviation, RSD, n = 5) and selectivity toward glucose, even in the presence of possible interfering species. In continuation with the previous work, Buk et al. [218] developed an electrochemical biosensor from the system CDQs/AuNPs and gold microdisk array electrodes. The biosensor presented an improved analytical performance and a sensitivity of 626.06 mA mM−1 ­cm−2 with a linear range from 0.16 mM to 4.32 mM. AuNPs coated with horseradish peroxidase (HRP) have been used to obtain an electrochemical aptasensor for the detection of kanamycin, based on streptavidin–biotin supramolecular interactions (see Fig. 7). After a biotinylated aptamer of kanamycin interacts with its complementary oligonucleotide strand modifying the electrode, NPs coated with HRP and streptavidin are added to probe the formation of the double-stranded DNA. Methylene blue is intercalated in the helical structure obtained as electron mediator [219]. Another attractive application of AuNPs is their compatibility to conjugate antigenic carbohydrates [220]. Zhang et  al. [221] described the design of a thioninebridged multiwalled carbon nanotube (MWCNT)/AuNP composite to detect a glycan on living cancer cells using enzyme catalysis. In such work, thionine allowed negatively charged AuNPs to bind to the MWCNT surface. This biosensing platform was used to quantify the amount of mannose at the surface of the cells, which corresponded to 3.39 × ­1010 molecules per human liver cancer cells, and 1.84 × 1­ 010 molecules for prostate cancer cells. Compared to other reported methods, this biosensor provided a useful protocol for the glycan assay, facilitating early medical diagnosis. In this regard, the field of biosensors has achieved remarkable advances in the detection of cancer biomarkers [222]. Recently, Nguyen et  al. [223] demonstrated a simple microfluidic device for the capture of A549 human lung adenocarcinoma

13

114

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12

cells using a specific aptamer and the SAM of AuNPs. Lima et  al. [224] have recently prepared the first electrochemical immunosensor to detect dehydroepiandrosterone sulfate (DHEAS) by electrochemical impedance spectroscopy (EIS), in order to provide an alternative method for the early diagnosis of pediatric adrenocortical carcinoma (pACC). The fabrication of this immunosensor consisted in the modification of an oxidized glassy carbon electrode (GCE) with arginine-functionalized AuNPs (AuNPs-ARG) and anti-DHEA IgM antibodies (ox-GCE/AuNPs-ARG/ IgM). The AuNP-based immunosensor showed good sensitivity, accuracy, stability, selectivity, and feasibility to detect DHEAS with a linear range from 10.0 to 110.0 µg ­dL−1, with a LOD of 7.4 µg ­dL−1. 6.2 Electrochemical Applications of Bioconjugates Based on AgNPs AgNPs are also important nanomaterials in electroanalysis due to their physicochemical properties. Many efforts have been made to improve the analytical methods that allow quantification of different biomarkers, metabolites and infectious agents based on AgNPs or nanocomposites [225]. Abbaspour and co-workers described a highly selective sandwich immunosensor based on a dual-aptamer for the detection of Staphylococcus aureus [226]. In this latter study, the authors first immobilized a biotinylated anti-S. aureus aptamer on streptavidin-coated magnetic beads via biotin-streptavidin affinity reaction (see section Iron oxide NPs and Fig. 7). Subsequently, the AgNPs conjugated with antiS. aureus aptamer (Apt-AgNP) were incorporated, thus completing the sandwich design. Here, the AgNPs were assembled to the aptamer through their thiol groups (see section Gold and silver nanoparticles and Fig. 5). The biosensor presented high sensitivity and an extended dynamic range from 10 to 1 × ­106 CFU/mL with a low detection limit of 1.0  CFU/mL (S/N = 3). AgNPs are also useful for the detection of pharmaceutical drugs. In this regard, AgNPs play an important role as electrode modifiers due to their ability to increase the conductivity in the biosensor [225]. Ashrafi et  al. [227] developed a novel, unique and sensitive biosensor to determine benzodiazepines, i.e., alprazolam, chlordiazepoxide, diazepam, oxazepam, and clonazepam. To prepare the biosensor, the authors synthesized a nano-ink based on AgNPs plus N-doped graphene quantum dots (Ag/N-GQD) and then electrodeposited it at the surface of a gold electrode modified with chitosan (CS), using LBL strategy. The obtained CS-Ag/N-GQD film combined the advantages of CS, and Ag/N-GQD with excellent electrical conductivity, biocompatibility and abundant active sites for the electro-oxidation of the species of interest within standard and plasma samples. Another fine example has been reported by Roushani and Shahdost-fard [228]. These authors developed a selective electrochemical aptasensor for the ultrasensitive detection of cocaine. The aptasensor was constructed by the covalent immobilization of aptamer-functionalized AgNPs as biorecognition element, on top of a nanocomposite made of MWCNTs, an ionic liquid (IL) and CS (MWCNTs/IL/ CS). Riboflavin was used as the redox probe in the electrochemical aptasensor for the diagnosis of the target. The biosensor showed high sensitivity, specificity, and

Reprinted from the journal

115

13



Topics in Current Chemistry (2020) 378:12

selectivity with a detection limit of 0.15 nM. Likewise, Aftab et al. [229] recently published a sensitive voltammetric nanosensor for the detection of Rilpivirine (RIL) based on amine group functionalized MWCNT, CQDs and AgNPs (CQDs/NH2fMWCNT/AgNPs). The nanosensor allowed detection of RIL in biological samples with LOD of 1.79 × 10−10 M, 4.47 × 10−10 M in serum sample and 5.26 × 10−10 M, 8.23 × 10−10 M. There is a significant interest in the early detection of different types of cancer. Meng et al. [230] developed an electrochemical biosensor for the sensitive analysis of prostate-specific antigen (PSA) based on a specific peptide as a molecular recognition element. The peptide was immobilized onto a gold electrode surface. In the absence of PSA, GO is immobilized on the peptide-modified electrode. In turn, it allows the silver ions to be absorbed onto GO through electrostatic interactions. GO acts as a reducing and dispersing agent, facilitating the formation of AgNPs, which contributes to better electronic transport and hence to the analytical signal. In the presence of PSA, the peptide is specifically recognized and cleaved at the surface of the electrode. The biosensor displayed high sensitivity, selectivity and a linear range from 5 to 2 × 104 pg/mL, with a LOD of 0.33 pg/mL. Elhakim et al. [231] designed a novel sensitive electrochemical sensor for microribonucleic acid (microRNA) detection in normal serum samples, hepatocellular carcinoma patients and human liver cancer cells. The biosensor constructed consisted in a carbon paste (CP) decorated with AgNPs and extracted propolis (bee glue). The porous structure of propolis allowed the AgNPs to be adsorbed, increasing the surface area of the sensor, facilitating the charge transfer process. The microRNA was also immobilized on the surface of the electrode through trapping with propoleos. The biosensor presented high selectivity and sensitivity, and a very low detection limit of ­10−3 femtomolar. Chen and co-authors reported the fabrication of inexpensive and flexible electronic and electrochemical sensors for a wide range of biochemical and biomedical applications [232]. The electrochemical biosensor was fabricated by a simple method involving wax patterning on plastic, hand painting of AgNPs to lay down a conducting layer, and simple drop casting of carbon nanotubes (CNT) to improve the electrochemical performance of the sensor. The nanosensor was applied for the amperometric detection of carcinoembryonic antigen (CEA) by monitoring an electroactive product released from a magnetic-bead based immunoassay. The limit of detection for CEA found to be 0.46 ng/mL, which is 10 times lower than the clinical cutoff value. The results of this work show that the improvement in electrochemical performance can be achieved by simple drop casting of CNT onto working AgNP electrodes. 6.3 Electrochemical Applications of Bioconjugates Based on IONPs MNPs are also used widely in the development of biosensors. In this sense, magnetic IONPs have been studied intensively due to their strong magnetic properties, electrocatalytic activity, biocompatibility with biomolecules, inexpensive synthesis and low toxicity [108, 130]. For example, electrocatalytic processes on the electrode

13

116

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



surfaces can be switched by means of functionalized IONPs in the presence of an external magnet [233]. In addition, the F ­ e3O4 NPs possesses peroxidase-like activity that could catalyze the electrochemical oxidation–reduction of ­H2O2, similar to HRP [234]. It is worth noting that both the oxidation reaction and the reduction of ­H2O2, depending on the applied potential, can be used as a transduction reaction in a biosensor [235]. Tian et  al. [236] have designed an ultrasensitive electrochemical cytosensor to detect MCF-7 CTCs—a tumor marker in human breast carcinoma. The electrochemical cytosensor was developed based on magnetic field-induced, targeted separation, and enrichment, and RGO/molybdenum disulfide (RGO/MoS2) composites and ­Fe3O4 NPs with a synergistic effect on the catalysis of H ­ 2O2. The proposed cytosensor exhibited a linear range from 15 to 45 cells ­mL−1 with a lower detection limit of 6 cells ­mL−1 for MCF-7 detection. Pakapongpan and Poo-arporn [211] describe a novel strategy for obtaining an enzyme biosensor based on direct electrochemistry. In this case, nanomaterials such as RGO and ­Fe3O4 NPs were used to increase the specific area of the electrode, to create a favorable environment to immobilize the enzyme GOx and to facilitate electron transfer between the enzyme and the surface of the electrode. Initially, the authors obtained the RGO-Fe3O4 nanocomposite by means of covalent bonding through the coupling agent, EDC, and NHS (see Scheme  1), while for immobilization of the enzyme they used electrostatic interactions since GOx is negatively charged at pH 7 and on the surface of the F ­ e3O4 NPs there is a positive environment of amino groups. Recently, ­Fe3O4/graphene composites have attracted great interest in the manufacture of biosensors. Teymourian et al. [237] established a novel label-free nanocomposite, using an indicator-free strategy of electrochemical DNA sensor based on ­Fe3O4 NPs/RGO ­(Fe3O4/RGO) nanocomposite. In the functioning of the sensor, ­Fe3O4/RGO nanocomposite was used as a substrate to immobilize probe DNA and hybridization with the target sequence to form dsDNA, which achieved a better analytical signal through measuring changes in the differential pulse voltammetric (DPV) peak current of the underlying Fe(II)/Fe(III) redox system. ­ e3O4, Tufa et al. [238] described an electrochemical immunosensor using GQD, F and AgNPs for the detection of Mycobacterium tuberculosis. In this device, the synergic effect of three nanomaterials is used to achieve better electrochemical performance. A core–shell ­Fe3O4@Ag/GQDs nanotriplex was synthesised for immobilization of Ab1, and by means of the Ab2-AuNPs conjugate as the label for detection of the culture filtrate protein (CFP-10). ­Fe3O4 NPs were used to improve mass transport and increase the surface to volume ratio, while AgNPs enhanced electrical conductivity, helping prevent ­Fe3O4 NPs from aggregating and helping GQD load more of the anti-CFP-10 antibody onto the electrode. The immunosensor showed a wide linear range (0.005–500 μg/mL) with a limit of detection (signal/noise  =  3) reaching 0.33 ng/mL, resulting in a reliable and robust performance with high selectivity toward CFP-10. An electrochemical biosensor based on ­Fe3O4 NPs–PVA nanocomposite for glucose detection has also been reported by Sanaeifar et al. [207]. The polyvinyl alcohol-Fe3O4 nanocomposite acted as a modifier of the surface of an Sn electrode and Reprinted from the journal

117

13



Topics in Current Chemistry (2020) 378:12

allowed GOx to be immobilized via physical adsorption. Also, F ­ e3O4 NPs dispersed in the PVA matrix promoted electron transfer between the enzyme and electrode. The GOx/PVA-Fe3O4/Sn bioelectrode detected glucose in the range from 5 × 10−3 to 30 mM with a sensitivity of 9.36 μA ­mM−1, and exhibited a lower detection limit of 8 μM at a signal-to-noise ratio of 3. Shamsipur et  al. [239] described a sandwich-type immunoassay for the detection of human epidermal growth factor receptor 2 (HER2)—a key prognostic tumor marker allowing diagnosis of breast cancer at early stages. The electrochemical immunosensor employed the synergistic effect of a nanomagnetic platform and a bioconjugate for the determination of HER2. ­Fe3O4 nanoparticles were synthesized by a simple chemical co-precipitation method. The magnetite nanoparticles coated with 3-aminopropyltrimethoxysilane and conjugated to the antibody were then immobilized on a GCE. The biosensor showed a linear relationship between the DPV stripping signal of silver and the logarithm of HER2 concentrations was obtained in the range of 5.0 × 10−4–50.0 ng/mL (R2 = 0.9906) with a detection limit of 2.0 × 10−5 ng/mL. Kumar et  al. [240] presented an easy method for the fabrication of an electrochemical paper biosensor to detect CEA biomarker using ­Fe3O4 NPs decorated poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS). Whatman filter paper (WP) was chosen for preparation of the paper electrode. The incorporation of F ­ e3O4 NPs into PEDOT:PSS/WP improved electrochemical performance and signal stability. The nanostructured paper-based electrode was used to immobilize anti-CEA monoclonal antibodies. The proposed immunoelectrode exhibited efficient linear detection range (4–25 ng/mL), high sensitivity (10.2 µA ­ng−1 mL cm−2) and long shelf life (34 days). Chikhaliwala et  al. [241] constructed an electrochemical immunoassay for the simultaneous determination of alpha-fetoprotein (AFP) and glypican-3 (GPC-3), both constitute important biomarkers for early detection of hepatocellular carcinoma (HCC). The GCE was modified through drop-casting with F ­ e3O4 NPs decorated ­ e3O4 NPs facilitated with hyperbranched amino functionalized dendrimers. The F magnetic separation of analytes as well as enhanced the electroanalytical effect in synergy with a large surface area of dendrimers. This sensor exhibited a linear range from 0.02 to 10  ng/mL for both biomarkers, and a low detection limit of 50 and 70 pg/mL, respectively.

7 Conclusions This review is dedicated to gold, silver, and iron oxide nanoparticles, their main physical and chemical characteristics, synthetic routes, and surface modification with ligands and biomolecules aimed at electrochemical applications. We have provided an overview of the multiple designs and architectures, along with experimental procedures to obtain functional bioconjugates and their use for electrochemical applications. A generalized view of the contents of this review shows that no one bionanoconjugation strategy is better than others. Indeed, selection of the methodology to

13

118

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



follow depends strongly on the proposed application, but the outcome results are difficult to predict a priori. However, some general guidelines may be followed. For certain applications, the orientation of the biomolecule on a support is not essential; when purifying a biomolecule over a (nanostructured) inorganic support, electrostatic interactions may be used for retention of the biomolecule, which may be further liberated after a change of pH or ionic strength. Conversely, antigen–antibody and enzyme-based systems often require that the biomolecule be orientated optimally so as not to block their recognition sites. Regarding the chemical modifications used to obtain nanobioconjugates, current methods are effective but the range of experimental conditions in which they works appropriately are essentially narrow, i.e., temperature, incubation and reaction times, reagents, pH and ionic strength are some of the factors to consider very carefully. It is also worth highlighting that, although nanoparticles modified with biomolecules have demonstrated high efficiency in diverse fields, many of the current systems exist only at the laboratory scale. With an eye on future developments in this field, we foresee scaling-up to gram-scale production as the next challenge in order to widen the use of nanobioconjugates in point-of-care diagnostic devices, thus increasing their impact in everyday life. Undeniably, this is a dynamic research field that will continue to grow exponentially in the near future.

References 1. Vial S, Reis RL, Oliveira JM (2017) Recent advances using gold nanoparticles as a promising multimodal tool for tissue engineering and regenerative medicine. Curr Opin Solid State Mater Sci 21:92–112. https​://doi.org/10.1016/j.cossm​s.2016.03.006 2. Sanad M, Shalan A, Bazid S et al (2019) A graphene gold nanocomposite-based 5-FU drug and the enhancement of the MCF-7 cell line treatment. RSC Adv 9:31021–31029. https​://doi.org/10.1039/ C9RA0​5669F​ 3. Ahsan MA, Jabbari V, Imam MA et al (2020) Nanoscale nickel metal organic framework decorated over graphene oxide and carbon nanotubes for water remediation. Sci Total Environ 698:134214. https​://doi.org/10.1016/j.scito​tenv.2019.13421​4 4. Ahsan MA, Jabbari V, Islam MT et  al (2019) Sustainable synthesis and remarkable adsorption capacity of MOF/graphene oxide and MOF/CNT based hybrid nanocomposites for the removal of Bisphenol A from water. Sci Total Environ 673:306–317. https​://doi.org/10.1016/j.scito​ tenv.2019.03.219 5. Ahn J, Choi Y, Lee AR et al (2016) A duplex DNA-gold nanoparticle probe composed as a colorimetric biosensor for sequence-specific DNA-binding proteins. Analyst 141:2040–2045. https​://doi. org/10.1039/c6an0​0033a​ 6. Chen M, Kang H, Gong Y et al (2015) Bacterial cellulose supported gold nanoparticles with excellent catalytic properties. ACS Appl Mater Interfaces 7:21717–21726. https​://doi.org/10.1021/ acsam​i.5b071​50 7. Rodriguez-Padron D, Santiago ARP, Balu AM, Luque R (2018) Synthesis of high valuable N-heterocycles via catalytic conversion of levulinic acid. Front Chem 6:662. https​://doi.org/10.3389/ fchem​.2018.00662​ 8. Ahsan MA, Jabbari V, El-Gendy AA et  al (2019) Ultrafast catalytic reduction of environmental pollutants in water via MOF-derived magnetic Ni and Cu nanoparticles encapsulated in porous carbon. Appl Surf Sci 497:143608. https​://doi.org/10.1016/j.apsus​c.2019.14360​8 9. Ahsan MA, Deemer E, Fernandez-Delgado O et al (2019) Fe nanoparticles encapsulated in MOFderived carbon for the reduction of 4-nitrophenol and methyl orange in water. Catal Commun 130:105753. https​://doi.org/10.1016/j.catco​m.2019.10575​3 Reprinted from the journal

119

13



Topics in Current Chemistry (2020) 378:12

10. Ahsan MA, Fernandez-Delgado O, Deemer E et al (2019) Carbonization of Co-BDC MOF results in magnetic C@Co nanoparticles that catalyze the reduction of methyl orange and 4-nitrophenol in water. J Mol Liq 290:111059. https​://doi.org/10.1016/j.molli​q.2019.11105​9 11. Rodríguez-Padrón D, Puente-Santiago AR, Balu AM et al (2019) Environmental catalysis: present and future. ChemCatChem 11:18–38. https​://doi.org/10.1002/cctc.20180​1248 12. Cova CM, Zuliani A, Puente Santiago AR et al (2018) Microwave-assisted preparation of Ag/Ag2S carbon hybrid structures from pig bristles as efficient HER catalysts. J Mater Chem A 6:21516– 21523. https​://doi.org/10.1039/c8ta0​6417b​ 13. Franco A, Cebrián-García S, Rodríguez-Padrón D et al (2018) Encapsulated laccases as effective electrocatalysts for oxygen reduction reactions. ACS Sustain Chem Eng 6:11058–11062. https​:// doi.org/10.1021/acssu​schem​eng.8b025​29 14. Wongkaew N, Simsek M, Griesche C, Baeumner AJ (2019) Functional nanomaterials and nanostructures enhancing electrochemical biosensors and lab-on-a-chip performances: recent progress, applications, and future perspective. Chem Rev 119:120–194. https​://doi.org/10.1021/acs.chemr​ ev.8b001​72 15. Lu M, Zhu H, Bazuin CG et al (2019) Polymer-templated gold nanoparticles on optical fibers for enhanced-sensitivity localized surface plasmon resonance biosensors. ACS Sensors 4:613–622. https​://doi.org/10.1021/acsse​nsors​.8b013​72 16. Villalonga ML, Borisova B, Arenas CB et  al (2019) Disposable electrochemical biosensors for Brettanomyces bruxellensis and total yeast content in wine based on core–shell magnetic nanoparticles. Sensors Actuators, B Chem 279:15–21. https​://doi.org/10.1016/j.snb.2018.09.092 17. Sebastian Maria, Aravind A, Mathew B (2018) Green silver nanoparticles based dual sensor for toxic Hg(II) ions. Nanotechnology 29(35):355502 18. Prakash A, Pathrose BP, Mathew S et  al (2018) Variations in thermo-optical properties of neutral red dye with laser ablated gold nanoparticles. Opt Mater (Amst) 79:237–242. https​://doi. org/10.1016/j.optma​t.2018.03.044 19. dos Santos Courrol D, Regina Borges Lopes C, da Silva Cordeiro T et al (2018) Optical properties and antimicrobial effects of silver nanoparticles synthesized by femtosecond laser photoreduction. Opt Laser Technol 103:233–238. https​://doi.org/10.1016/j.optla​stec.2018.01.044 20. Syrek K, Grudzień J, Sennik-Kubiec A et  al (2019) Anodic titanium oxide layers modified with gold, silver, and copper nanoparticles. J Nanomater 2019:1–10. https​://doi.org/10.1155/2019/92087​ 34 21. Illés E, Szekeres M, Tóth IY et  al (2018) Multifunctional PEG-carboxylate copolymer coated superparamagnetic iron oxide nanoparticles for biomedical application. J Magn Magn Mater 122:710–720. https​://doi.org/10.1590/1980-5373-MR-2018-0094 22. Aghazadeh M, Karimzadeh I, Maragheh MG, Ganjali MR (2018) Enhancing the supercapacitive and superparamagnetic performances of iron oxide nanoparticles through yttrium cations electrochemical doping. Mater Res 29:2291–2300. https​://doi.org/10.1590/1980-5373-MR-2018-0094 23. Ponnaiah SK, Periakaruppan P, Vellaichamy B (2018) New electrochemical sensor based on a silver-doped iron oxide nanocomposite coupled with polyaniline and its sensing application for picomolar-level detection of uric acid in human blood and urine samples. J Phys Chem B 122:3037– 3046. https​://doi.org/10.1021/acs.jpcb.7b115​04 24. Seifan M, Ebrahiminezhad A, Ghasemi Y et al (2018) Amine-modified magnetic iron oxide nanoparticle as a promising carrier for application in bio self-healing concrete. Appl Microbiol Biotechnol 102:175–184. https​://doi.org/10.1007/s0025​3-017-8611-z 25. Khramtsov P, Kropaneva M, Byzov I et  al (2019) Conjugation of carbon coated-iron nanoparticles with biomolecules for NMR-based assay. Coll Surf B Biointerfaces 176:256–264. https​://doi. org/10.1016/j.colsu​rfb.2019.01.009 26. Santos JJ, Leal J, Dias LAP et al (2018) Bovine serum albumin conjugated gold-198 nanoparticles as model to evaluate damage caused by ionizing radiation to biomolecules. ACS Appl Nano Mater 1:5062–5070. https​://doi.org/10.1021/acsan​m.8b011​74 27. Qiao R, Esser L, Fu C et al (2018) Bioconjugation and fluorescence labeling of iron oxide nanoparticles grafted with bromomaleimide-terminal polymers. Biomacromol 19:4423–4429. https​://doi. org/10.1021/acs.bioma​c.8b012​82 28. Manivannan S, Seo Y, Kang DK, Kim K (2018) Colorimetric and optical Hg(ii) ion sensor developed with conjugates of M13-bacteriophage and silver nanoparticles. New J Chem 42:20007– 20014. https​://doi.org/10.1039/c8nj0​4496a​

13

120

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12 29. Chen H, Hu H, Tao C et al (2019) Self-assembled Au@Fe core/satellite magnetic nanoparticles for versatile biomolecule functionalization. ACS Appl Mater Interfaces 11:23858–23869. https​://doi. org/10.1021/acsam​i.9b055​44 30. Walper SA, Turner KB, Medintz IL (2015) Enzymatic bioconjugation of nanoparticles: developing specificity and control. Curr Opin Biotechnol 34:232–241. https​://doi.org/10.1016/j.copbi​ o.2015.04.003 31. Kairdolf BA, Qian X, Nie S (2017) Bioconjugated nanoparticles for biosensing, in vivo imaging, and medical diagnostics. Anal Chem 89:1015–1031. https​://doi.org/10.1021/acs.analc​hem.6b048​ 73 32. Austin LA, MacKey MA, Dreaden EC, El-Sayed MA (2014) The optical, photothermal, and facile surface chemical properties of gold and silver nanoparticles in biodiagnostics, therapy, and drug delivery. Arch Toxicol 88:1391–1417. https​://doi.org/10.1007/s0020​4-014-1245-3 33. Roduner E (2006) Size matters: why nanomaterials are different. Chem Soc Rev 35:583–592. https​ ://doi.org/10.1039/b5021​42c 34. Manoharan H, Kalita P, Gupta S, Sai VVR (2019) Plasmonic biosensors for bacterial endotoxin detection on biomimetic C-18 supported fiber optic probes. Biosens Bioelectron 129:79–86. https​ ://doi.org/10.1016/j.bios.2018.12.045 35. Amendola V, Pilot R, Frasconi M et al (2017) Surface plasmon resonance in gold nanoparticles: a review. J Phys Condens Matter 29:203002 36. Kuisma M, Sakko A, Rossi TP et al (2015) Localized surface plasmon resonance in silver nanoparticles: atomistic first-principles time-dependent density-functional theory calculations. Phys Rev B Condens Matter Mater Phys 91:1–8. https​://doi.org/10.1103/PhysR​evB.91.11543​1 37. Mie G (1908) Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen. Ann Phys 330:377–445. https​://doi.org/10.1002/andp.19083​30030​2 38. Huang X, El-Sayed MA (2010) Gold nanoparticles: optical properties and implementations in cancer diagnosis and photothermal therapy. J Adv Res 1:13–28. https​://doi.org/10.1016/j. jare.2010.02.002 39. Amendola V, Meneghetti M (2009) Size evaluation of gold nanoparticles by UV-Vis spectroscopy. J Phys Chem C 113:4277–4285. https​://doi.org/10.1021/jp808​2425 40. Jain PK, Lee KS, El-sayed IH, El-sayed MA (2006) Calculated absorption and scattering properties of gold nanoparticles of different size, shape, and composition: applications in biological imaging and biomedicine. J Phys Chem B 110:7238–7248. https​://doi.org/10.1021/jp057​170o 41. Chang X, Wang Y-F, Zhang X et  al (2019) Iridium size effects in localized surface plasmonenhanced diamond UV photodetectors. Appl Surf Sci 487:674–677. https​://doi.org/10.1016/j.apsus​ c.2019.04.268 42. Chen Y, Chen L, Wu Y, Di J (2019) Highly sensitive determination of dopamine based on the aggregation of small-sized gold nanoparticles. Microchem J 147:955–961. https​://doi. org/10.1016/j.micro​c.2019.04.025 43. Karami P, Khoshsafar H, Johari-Ahar M et al (2019) Colorimetric immunosensor for determination of prostate specific antigen using surface plasmon resonance band of colloidal triangular shape gold nanoparticles. Spectrochim Acta A 222:117218. https​://doi.org/10.1016/j.saa.2019.11721​8 44. Dissanayake NM, Arachchilage JS, Samuels TA, Obare SO (2019) Highly sensitive plasmonic metal nanoparticle-based sensors for the detection of organophosphorus pesticides. Talanta 200:218–227. https​://doi.org/10.1016/j.talan​ta.2019.03.042 45. Ke S, Kan C, Ni Y et al (2019) Construction of silica-encapsulated gold-silver core-shell nanorod: atomic facets enrichment and plasmon enhanced catalytic activity with high stability and reusability. Mater Des 177:107837. https​://doi.org/10.1016/j.matde​s.2019.10783​7 46. Mizuno A, Ono A (2019) Static and dynamic tuning of surface plasmon resonance by controlling interparticle distance in arrays of Au nanoparticles. Appl Surf Sci 480:846–850. https​://doi. org/10.1016/j.apsus​c.2019.03.058 47. Kwon NK, Lee TK, Kwak SK, Kim SY (2017) Aggregation-driven controllable plasmonic transition of silica-coated gold nanoparticles with temperature-dependent polymer-nanoparticle interactions for potential applications in optoelectronic devices. ACS Appl Mater Interfaces 9:39688– 39698. https​://doi.org/10.1021/acsam​i.7b131​23 48. Kim HM, Hong Jeong D, Lee HY et  al (2019) Improved stability of gold nanoparticles on the optical fiber and their application to refractive index sensor based on localized surface plasmon resonance. Opt Laser Technol 114:171–178. https​://doi.org/10.1016/j.optla​stec.2019.02.002

Reprinted from the journal

121

13



Topics in Current Chemistry (2020) 378:12

49. Topcu G, Guner T, Inci E, Demir MM (2019) Colorimetric and plasmonic pressure sensors based on polyacrylamide/Au nanoparticles. Sensors Actuat A-Phys 295:503–511. https​://doi. org/10.1016/j.sna.2019.06.038 50. Vishnoi R, Sharma K, Sharma GD, Singhal R (2019) Temperature induced surface plasmon resonance in Au/a-C nanocomposite thin film. Vacuum 167:40–46. https​://doi.org/10.1016/j.vacuu​ m.2019.05.031 51. Su C (2017) Environmental implications and applications of engineered nanoscale magnetite and its hybrid nanocomposites: a review of recent literature. J Hazard Mater 322:48–84. https​://doi. org/10.1016/j.jhazm​at.2016.06.060 52. Morrish AH (2001) The physical principles of magnetism. IEEE Magnetics Society. Wiley-IEEE, New York 53. Kolhatkar AG, Jamison AC, Litvinov D et al (2013) Tuning the magnetic properties of nanoparticles. Int J Mol Sci 14:15977–16009. https​://doi.org/10.3390/ijms1​40815​977 54. Leslie-Pelecky DL, Rieke RD (1996) Magnetic properties of nanostructured materials. Chem Mater 8:1770–1783. https​://doi.org/10.1021/cm960​077f 55. Soares PIP, Laia CAT, Carvalho A et al (2016) Iron oxide nanoparticles stabilized with a bilayer of oleic acid for magnetic hyperthermia and MRI applications. Appl Surf Sci 383:240–247. https​:// doi.org/10.1016/j.apsus​c.2016.04.181 56. Kandasamy G, Sudame A, Bhati P et  al (2018) Systematic investigations on heating effects of carboxyl-amine functionalized superparamagnetic iron oxide nanoparticles (SPIONs) based ferrofluids for in vitro cancer hyperthermia therapy. J Mol Liq 256:224–237. https​://doi.org/10.1016/j. molli​q.2018.02.029 57. Ebrahimisadr S, Aslibeiki B, Asadi R (2018) Magnetic hyperthermia properties of iron oxide nanoparticles: the effect of concentration. Phys C 549:119–121. https​://doi.org/10.1016/j.physc​ .2018.02.014 58. Cornell RM, Schwertmann U (2003) Crystal structure. In: The iron oxides: structures, properties, reactions, occurrences and uses, 2nd, completely revised and extended edition. Wiley-VCH, Weinheim, pp 9–33 59. Qin L, Zeng G, Lai C et al (2018) “Gold rush” in modern science: fabrication strategies and typical advanced applications of gold nanoparticles in sensing. Coord Chem Rev 359:1–31. https​://doi. org/10.1016/j.ccr.2018.01.006 60. Liu A, Wang G, Wang F, Zhang Y (2017) Gold nanostructures with near-infrared plasmonic resonance: synthesis and surface functionalization. Coord Chem Rev 336:28–42. https​://doi. org/10.1016/j.ccr.2016.12.019 61. Turkevich J, Stevenson PC, Hillier J (1951) A study of the nucleation and growth processes in the synthesis of colloidal gold. Discuss Faraday Soc 11:55–75. https​://doi.org/10.1039/DF951​11000​55 62. Bartosewicz B, Bujno K, Liszewska M et  al (2018) Effect of citrate substitution by various α-hydroxycarboxylate anions on properties of gold nanoparticles synthesized by Turkevich method. Coll Surf A Physicochem Eng Asp 549:25–33. https​://doi.org/10.1016/j.colsu​rfa.2018.03.073 63. Brust M, Walker M, Bethell D et  al (1994) Synthesis of thiol-derivatised gold nanoparticles in a two-phase liquid–liquid system. J Chem Soc Chem Commun 1994:801–802. https​://doi. org/10.1039/C3994​00008​01 64. Kim YJ, Park J, Jeong HS et al (2019) A seed-mediated growth of gold nanoparticles inside carbon nanotube fibers for fabrication of multifunctional nanohybrid fibers with enhanced mechanical and electrical properties. Nanoscale 11:5295–5303. https​://doi.org/10.1039/c8nr1​0446h​ 65. Song C, Li F, Guo X et al (2019) Gold nanostars for cancer cell-targeted SERS-imaging and NIR light-triggered plasmonic photothermal therapy (PPTT) in the first and second biological windows. J Mater Chem B 7:2001–2008. https​://doi.org/10.1039/c9tb0​0061e​ 66. Wiesner J, Wokaun A (1989) Anisometric gold colloids. Preparation, characterization, and optical properties. Chem Phys Lett 157:569–575. https​://doi.org/10.1016/S0009​-2614(89)87413​-5 67. Philip A, Ankudze B, Pakkanen TT (2018) Polyethylenimine-assisted seed-mediated synthesis of gold nanoparticles for surface-enhanced Raman scattering studies. Appl Surf Sci 444:243–252. https​://doi.org/10.1016/j.apsus​c.2018.03.042 68. Vágó A, Szakacs G, Sáfrán G et al (2016) One-step green synthesis of gold nanoparticles by mesophilic filamentous fungi. Chem Phys Lett 645:1–4. https​://doi.org/10.1016/j.cplet​t.2015.12.019 69. Vimalraj S, Ashokkumar T, Saravanan S (2018) Biogenic gold nanoparticles synthesis mediated by Mangifera indica seed aqueous extracts exhibits antibacterial, anticancer and anti-angiogenic properties. Biomed Pharmacother 105:440–448. https​://doi.org/10.1016/j.bioph​a.2018.05.151

13

122

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



70. Osonga FJ, Kariuki VM, Wambua VM et al (2019) Photochemical synthesis and catalytic applications of gold nanoplates fabricated using quercetin diphosphate macromolecules. ACS Omega 4:6511–6520. https​://doi.org/10.1021/acsom​ega.8b023​89 71. Alluhaybi HA, Ghoshal SK, Shamsuri WNW et al (2019) Pulsed laser ablation in liquid assisted growth of gold nanoparticles: evaluation of structural and optical features. Nano-Structures NanoObjects 19:100355. https​://doi.org/10.1016/j.nanos​o.2019.10035​5 72. Mitra C, Gummadidala PM, Afshinnia K et al (2017) Citrate-coated silver nanoparticles growthindependently inhibit aflatoxin synthesis in Aspergillus parasiticus. Environ Sci Technol 51:8085– 8093. https​://doi.org/10.1021/acs.est.7b012​30 73. Zhang W, Hu G, Zhang W et al (2019) A facile strategy for the synthesis of silver nanostructures with different morphologies. Mater Chem Phys 235:121629. https​://doi.org/10.1016/j.match​emphy​ s.2019.05.017 74. Onofre-Cordeiro NA, Silva YEO, Solidônio EG et al (2018) Agarose-silver particles films: effect of calcium ascorbate in nanoparticles synthesis and film properties. Int J Biol Macromol 119:701– 707. https​://doi.org/10.1016/j.ijbio​mac.2018.07.115 75. Kharissova OV, Dias HVR, Kharisov BI et  al (2013) The greener synthesis of nanoparticles. Trends Biotechnol 31:240–248. https​://doi.org/10.1016/j.tibte​ch.2013.01.003 76. Hernández-Morales L, Espinoza-Gómez H, Flores-López LZ et al (2019) Study of the green synthesis of silver nanoparticles using a natural extract of dark or white Salvia hispanica L. seeds and their antibacterial application. Appl Surf Sci 489:952–961. https​://doi.org/10.1016/j.apsus​ c.2019.06.031 77. Akter M, Sikder MT, Rahman MM et  al (2018) A systematic review on silver nanoparticlesinduced cytotoxicity: physicochemical properties and perspectives. J Adv Res 9:1–16. https​://doi. org/10.1016/j.jare.2017.10.008 78. Kumar V, Bano D, Mohan S et al (2016) Sunlight-induced green synthesis of silver nanoparticles using aqueous leaf extract of Polyalthia longifolia and its antioxidant activity. Mater Lett 181:371– 377. https​://doi.org/10.1016/j.matle​t.2016.05.097 79. Moodley JS, Krishna SBN, Pillay K et al (2018) Green synthesis of silver nanoparticles from Moringa oleifera leaf extracts and its antimicrobial potential. Adv Nat Sci Nanosci Nanotechnol 9:1–9. https​://doi.org/10.1088/2043-6254/aaabb​2 80. Singh H, Du J, Singh P, Yi TH (2018) Extracellular synthesis of silver nanoparticles by Pseudomonas sp. THG-LS1.4 and their antimicrobial application. J Pharm Anal 8:258–264. https​://doi. org/10.1016/j.jpha.2018.04.004 81. Zhang D, Zhang R, He X et al (2019) Facile fabrication of silver decorated polyarylene ether nitrile composited micro/nanospheres via microemulsion self-assembling. Compos Part B Eng 156:399– 405. https​://doi.org/10.1016/j.compo​sites​b.2018.08.046 82. Jahan I, Erci F, Isildak I (2019) Microwave-assisted green synthesis of non-cytotoxic silver nanoparticles using the aqueous extract of Rosa santana (rose) petals and their antimicrobial activity. Anal Lett 52:1860–1873. https​://doi.org/10.1080/00032​719.2019.15721​79 83. Rafique M, Rafique MS, Kalsoom U et al (2019) Laser ablation synthesis of silver nanoparticles in water and dependence on laser nature. Opt Quantum Electron 51:179. https​://doi.org/10.1007/ s1108​2-019-1902-0 84. Massart R (1981) Preparation of aqueous magnetic liquids in alkaline and acidic media. IEEE Trans Magn 17:1247–1248. https​://doi.org/10.1109/TMAG.1981.10611​88 85. Nikitin A, Fedorova M, Naumenko V et  al (2017) Synthesis, characterization and MRI application of magnetite water-soluble cubic nanoparticles. J Magn Magn Mater 441:6–13. https​://doi. org/10.1016/j.jmmm.2017.05.039 86. Unni M, Uhl AM, Savliwala S et al (2017) Thermal decomposition synthesis of iron oxide nanoparticles with diminished magnetic dead layer by controlled addition of oxygen. ACS Nano 11:2284–2303. https​://doi.org/10.1021/acsna​no.7b006​09 87. Wei Y, Zhang C, Chang Q et al (2017) Synthesis of monodisperse iron oxide nanoparticles: effect of temperature, time, solvent, and surfactant. Inorg Nano-Metal Chem 47:1375–1379. https​://doi. org/10.1080/24701​556.2017.12841​24 88. Lassenberger A, Grünewald TA, Van Oostrum PDJ et al (2017) Monodisperse iron oxide nanoparticles by thermal decomposition: elucidating particle formation by second-resolved in situ smallangle X-ray scattering. Chem Mater 29:4511–4522. https​://doi.org/10.1021/acs.chemm​ater.7b012​ 07

Reprinted from the journal

123

13



Topics in Current Chemistry (2020) 378:12

89. Jović Orsini N, Babić-Stojić B, Spasojević V et al (2018) Magnetic and power absorption measurements on iron oxide nanoparticles synthesized by thermal decomposition of Fe(acac)3. J Magn Magn Mater 449:286–296. https​://doi.org/10.1016/j.jmmm.2017.10.053 90. Jia X, Chen X, Liu Y et  al (2019) Hydrophilic F ­ e3O4 nanoparticles prepared by ferrocene as high-efficiency heterogeneous fenton catalyst for the degradation of methyl orange. Appl Organomet Chem. https​://doi.org/10.1002/aoc.4826 91. Gao L, Tang Y, Wang C et  al (2019) Highly-efficient amphiphilic magnetic nanocomposites based on a simple sol–gel modification for adsorption of phthalate esters. J Colloid Interface Sci 552:142–152. https​://doi.org/10.1016/j.jcis.2019.05.031 92. Kamura A, Idota N, Sugahara Y (2019) Nonaqueous synthesis of magnetite nanoparticles via oxidation of tetrachloroferrate anions by pyridine-N-oxide. Solid State Sci 92:81–88. https​:// doi.org/10.1016/j.solid​state​scien​ces.2018.10.018 93. Mandriota G, Di Corato R, Benedetti M et al (2019) Design and application of cisplatin-loaded magnetic nanoparticle clusters for smart chemotherapy. ACS Appl Mater Interfaces 11:1864– 1875. https​://doi.org/10.1021/acsam​i.8b187​17 94. Svetlichnyi VA, Shabalina AV, Lapin IN et al (2019) Comparative study of magnetite nanoparticles obtained by pulsed laser ablation in water and air. Appl Surf Sci 467:402–410. https​://doi. org/10.1016/j.apsus​c.2018.10.189 95. Arsalani S, Guidelli EJ, Araujo JFDF et al (2018) Green synthesis and surface modification of iron oxide nanoparticles with enhanced magnetization using natural rubber latex. ACS Sustain Chem Eng 6:13756–13765. https​://doi.org/10.1021/acssu​schem​eng.8b016​89 96. Sperling RA, Parak WJ (2010) Surface modification, functionalization and bioconjugation of colloidal Inorganic nanoparticles. Philos Trans R Soc A Math Phys Eng Sci 368:1333–1383. https​://doi.org/10.1098/rsta.2009.0273 97. Nosrati H, Salehiabar M, Davaran S et al (2015) New advances strategies for surface functionalization of iron oxide magnetic nano particles (IONPs). Res Chem Intermed 43:7423–7442. https​://doi.org/10.1007/s1116​4-017-3084-3 98. Bohórquez AC, Unni M, Belsare S et al (2018) Stability and mobility of magnetic nanoparticles in biological environments determined from dynamic magnetic susceptibility measurements. Bioconjug Chem 29:2793–2805. https​://doi.org/10.1021/acs.bioco​njche​m.8b004​19 99. Kluenker M, Mondeshki M, Nawaz Tahir M, Tremel W (2018) Monitoring thiol-ligand exchange on Au nanoparticle surfaces. Langmuir 34:1700–1710. https​://doi.org/10.1021/acs. langm​uir.7b040​15 100. Wang X, Wang X, Bai X et  al (2019) Nanoparticle ligand exchange and its effects at the nanoparticle-cell membrane interface. Nano Lett 19:8–18. https​://doi.org/10.1021/acs.nanol​ ett.8b026​38 101. Locardi F, Canepa E, Villa S et al (2018) Thermogravimetry and evolved gas analysis for the investigation of ligand-exchange reaction in thiol-functionalized gold nanoparticles. J Anal Appl Pyrolysis 132:11–18. https​://doi.org/10.1016/j.jaap.2018.03.023 102. Cho JS, Jang W, Park S et al (2018) Thermally stable propanethiol–ligand exchanged Ag nanoparticles for enhanced dispersion in perovskite solar cells via an effective incorporation method. J Ind Eng Chem 61:71–77. https​://doi.org/10.1016/j.jiec.2017.12.002 103. Okada S, Nakahara Y, Watanabe M et al (2019) Crystallite size increase of silver nanoparticles by ligand exchange and subsequent washing process with antisolvent. J Nanosci Nanotechnol 19:4565–4570. https​://doi.org/10.1166/jnn.2019.16361​ 104. Na HK, Kim H, Son JG et  al (2019) Facile synthesis and direct characterization of surfacecharge-controlled magnetic iron oxide nanoparticles and their role in gene transfection in human leukemic T cell. Appl Surf Sci 483:1069–1080. https​://doi.org/10.1016/j.apsus​c.2019.04.059 105. Pelaz B, Del Pino P, Maffre P et al (2015) Surface functionalization of nanoparticles with polyethylene glycol: effects on protein adsorption and cellular uptake. ACS Nano 9:6996–7008. https​://doi.org/10.1021/acsna​no.5b013​26 106. Zong J, Cobb SL, Cameron NR et al (2017) Peptide-functionalized gold nanoparticles: versatile biomaterials for diagnostic and therapeutic applications. Biomater Sci 13:1797–1808. https​:// doi.org/10.1016/j.nano.2017.02.010 107. Saha B, Songe P, Evers TH, Prins MWJ (2017) The influence of covalent immobilization conditions on antibody accessibility on nanoparticles. Analyst 142:4247–4256. https​://doi. org/10.1039/c7an0​1424d​

13

124

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



108. Fortes CCS, Daniel-da-Silva AL, Xavier AMRB, Tavares APM (2017) Optimization of enzyme immobilization on functionalized magnetic nanoparticles for laccase biocatalytic reactions. Chem Eng Process Process Intensif 117:1–8. https​://doi.org/10.1016/j.cep.2017.03.009 109. Poorakbar E, Shafiee A, Saboury AA et al (2018) Synthesis of magnetic gold mesoporous silica nanoparticles core shell for cellulase enzyme immobilization: improvement of enzymatic activity and thermal stability. Process Biochem 71:92–100. https​://doi.org/10.1016/j.procb​io.2018.05.012 110. Badgujar KC, Bhanage BM (2017) Investigation of deactivation thermodynamics of lipase immobilized on polymeric carrier. Bioprocess Biosyst Eng 40:741–757. https​://doi.org/10.1007/s0044​ 9-017-1740-z 111. Nadar SS, Rathod VK (2016) Magnetic macromolecular cross linked enzyme aggregates (CLEAs) of glucoamylase. Enzyme Microb Technol 83:78–87. https​://doi.org/10.1016/j.enzmi​ ctec.2015.10.009 112. Facin BR, Valério A, Bresolin D et al (2018) Improving reuse cycles of Thermomyces lanuginosus lipase (NS-40116) by immobilization in flexible polyurethane. Biocatal Biotransf 36:372–380. https​://doi.org/10.1080/10242​422.2018.14588​42 113. Leidner A, Bauer J, Ebrahimi Khonachah M et  al (2019) Oriented immobilization of a delicate glucose-sensing protein on silica nanoparticles. Biomaterials 190:76–85. https​://doi.org/10.1016/j. bioma​teria​ls.2018.10.035 114. Nicolás P, Lassalle V, Ferreira ML (2018) Immobilization of CALB on lysine-modified magnetic nanoparticles: influence of the immobilization protocol. Bioprocess Biosyst Eng 41:171–184. https​ ://doi.org/10.1007/s0044​9-017-1855-2 115. Liu DM, Chen J, Shi YP (2018) Tyrosinase immobilization on aminated magnetic nanoparticles by physical adsorption combined with covalent crosslinking with improved catalytic activity, reusability and storage stability. Anal Chim Acta 1006:90–98. https​://doi.org/10.1016/j.aca.2017.12.022 116. Tabarzad M, Sharafi Z, Javidi J (2018) Covalent immobilization of coagulation factor VIII on magnetic nanoparticles for aptamer development. J Appl Biomater Funct Mater 16:161–170. https​:// doi.org/10.1177/22808​00018​76504​6 117. Waifalkar PP, Parit SB, Chougale AD et al (2016) Immobilization of invertase on chitosan coated γ-Fe2O3 magnetic nanoparticles to facilitate magnetic separation. J Colloid Interface Sci 482:159– 164. https​://doi.org/10.1016/j.jcis.2016.07.082 118. Ortiz E, Gallay P, Galicia L et  al (2019) Nanoarchitectures based on multi-walled carbon nanotubes non-covalently functionalized with Concanavalin A: a new building-block with supramolecular recognition properties for the development of electrochemical biosensors. Sensors Actuators, B Chem 292:254–262. https​://doi.org/10.1016/j.snb.2019.04.114 119. Kadu K, Ghosh G, Panicker L et al (2019) Role of surface charges on interaction of rod-shaped magnetic hydroxyapatite nanoparticles with protein. Colloids Surfaces B Biointerfaces 177:362– 369. https​://doi.org/10.1016/j.colsu​rfb.2019.02.021 120. Li D, Huang X, Wu Y et al (2016) Preparation of pH-responsive mesoporous hydroxyapatite nanoparticles for intracellular controlled release of an anticancer drug. Biomater Sci 4:272–280. https​:// doi.org/10.1039/c5bm0​0228a​ 121. Satzer P, Svec F, Sekot G, Jungbauer A (2016) Protein adsorption onto nanoparticles induces conformational changes: particle size dependency, kinetics, and mechanisms. Eng Life Sci 16:238– 246. https​://doi.org/10.1002/elsc.20150​0059 122. Martínez-Carmona M, Lozano D, Colilla M, Vallet-Regí M (2018) Lectin-conjugated pH-responsive mesoporous silica nanoparticles for targeted bone cancer treatment. Acta Biomater 65:393– 404. https​://doi.org/10.1016/j.actbi​o.2017.11.007 123. Koshkina O, Lang T, Thiermann R et al (2015) Temperature-triggered protein adsorption on polymer-coated nanoparticles in serum. Langmuir 31:8873–8881. https​://doi.org/10.1021/acs.langm​ uir.5b005​37 124. Cortés FB, Montoya T, Acevedo S et al (2018) Adsorption-desorption of n–C7 asphaltenes over micro- and nanoparticles of silica and its impact on wettability alteration. CTF Cienc Tecnol Futuro 6:89–106. https​://doi.org/10.29047​/01225​383.06 125. Cao Y, Wen L, Svec F et  al (2016) Magnetic AuNP@Fe3O4 nanoparticles as reusable carriers for reversible enzyme immobilization. Chem Eng J 286:272–281. https​://doi.org/10.1016/j. cej.2015.10.075 126. Meissner J, Prause A, Bharti B, Findenegg GH (2015) Characterization of protein adsorption onto silica nanoparticles: influence of pH and ionic strength. Colloid Polym Sci 293:3381–3391. https​:// doi.org/10.1007/s0039​6-015-3754-x Reprinted from the journal

125

13

Topics in Current Chemistry (2020) 378:12 127. Angioletti-Uberti S, Ballauff M, Dzubiella J (2018) Competitive adsorption of multiple proteins to nanoparticles: the Vroman effect revisited. Mol Phys 116:3154–3163. https​://doi. org/10.1080/00268​976.2018.14670​56 128. Liu J, Liu Y, Jin D et al (2019) Immobilization of trypsin onto large-pore mesoporous silica and optimization enzyme activity via response surface methodology. Solid State Sci 89:15–24. https​:// doi.org/10.1016/j.solid​state​scien​ces.2018.12.014 129. An GS, Chae DH, Hur JU et al (2018) Hollow-structured ­Fe3O4 @SiO2 nanoparticles: novel synthesis and enhanced adsorbents for purification of plasmid DNA. Ceram Int 44:18791–18795. https​ ://doi.org/10.1016/j.ceram​int.2018.07.111 130. Sosa-Acosta JR, Silva JA, Fernández-Izquierdo L et al (2018) Iron Oxide Nanoparticles (IONPs) with potential applications in plasmid DNA isolation. Coll Surf A Physicochem Eng Asp 545:167– 178. https​://doi.org/10.1016/j.colsu​rfa.2018.02.062 131. Wang X, Zhang S, Xu Y et al (2018) Ionic strength-responsive binding between nanoparticles and proteins. Langmuir 34:8264–8273. https​://doi.org/10.1021/acs.langm​uir.8b009​44 132. Wang Y, Ji X, Pang P et al (2018) Synthesis of Janus Au nanorods/polydivinylbenzene hybrid nanoparticles for chemo-photothermal therapy. J Mater Chem B 6:2481–2488. https​://doi.org/10.1039/ c8tb0​0233a​ 133. Sun Y, Zhang W, Wang B et al (2018) A supramolecular self-assembly strategy for upconversion nanoparticle bioconjugation. Chem Commun 54:3851–3854. https​://doi.org/10.1039/c8cc0​0708j​ 134. Lalaoui N, Rousselot-Pailley P, Robert V et al (2016) Direct electron transfer between a site-specific pyrene-modified laccase and carbon nanotube/gold nanoparticle supramolecular assemblies for bioelectrocatalytic dioxygen reduction. ACS Catal 6:1894–1900. https​://doi.org/10.1021/acsca​ tal.5b024​42 135. Manivannan S, Kim K (2016) Electrochemical biosensor utilizing supramolecular association of enzyme on sol−gel matrix embedded gold nanoparticles supported reduced graphene oxide−cyclodextrin nanocomposite. Electroanalysis 28:1608–1616. https​://doi.org/10.1002/elan.20150​1104 136. Dreesen L, Sartenaer Y, Humbert C et al (2004) Probing ligand-protein recognition with sum-frequency generation spectroscopy: the avidin-biocytin case. ChemPhysChem 5:1719–1725. https​:// doi.org/10.1002/cphc.20040​0213 137. Jain A, Barve A, Zhao Z et al (2017) Comparison of avidin, neutravidin, and streptavidin as nanocarriers for efficient siRNA delivery. Mol Pharm 14:1517–1527. https​://doi.org/10.1021/acs.molph​ armac​eut.6b009​33 138. Afsharan H, Khalilzadeh B, Tajalli H et  al (2016) A sandwich type immunosensor for ultrasensitive electrochemical quantification of p53 protein based on gold nanoparticles/graphene oxide. Electrochim Acta 188:153–164. https​://doi.org/10.1016/j.elect​acta.2015.11.133 139. D’Agata R, Palladino P, Spoto G (2017) Streptavidin-coated gold nanoparticles: critical role of oligonucleotides on stability and fractal aggregation. Beilstein J Nanotechnol 8:1–11. https​://doi. org/10.3762/bjnan​o.8.1 140. Ly TN, Park S, Park SJ (2016) Detection of HIV-1 antigen by quartz crystal microbalance using gold nanoparticles. Sensors Actuators, B Chem 237:452–458. https​://doi.org/10.1016/j. snb.2016.06.112 141. Taghdisi SM, Danesh NM, Beheshti HR et al (2016) A novel fluorescent aptasensor based on gold and silica nanoparticles for the ultrasensitive detection of ochratoxin A. Nanoscale 8:3439–3446. https​://doi.org/10.1039/c5nr0​8234j​ 142. Gao F, Jiang Y, Hu M et al (2016) Bienzymatic nanoreactors composed of chloroperoxidase–glucose oxidase on Au@Fe3O4 nanoparticles: dependence of catalytic performance on the bioarchitecture. Mater Des 111:414–420. https​://doi.org/10.1016/j.matde​s.2016.09.025 143. Xue SS, Tan CP, Chen MH et al (2017) Tumor-targeted supramolecular nanoparticles self-assembled from a ruthenium-β-cyclodextrin complex and an adamantane-functionalized peptide. Chem Commun 53:842–845. https​://doi.org/10.1039/c6cc0​8296c​ 144. Xue Q, Liu Z, Guo Y, Guo S (2015) Cyclodextrin functionalized graphene-gold nanoparticle hybrids with strong supramolecular capability for electrochemical thrombin aptasensor. Biosens Bioelectron 68:429–436. https​://doi.org/10.1016/j.bios.2015.01.025 145. Cheng JG, Zhang YM, Liu Y (2018) Supramolecular assembly of thiolated cyclodextrin and ferrocene derivative for controlled drug delivery. ChemNanoMat 4:758–763. https​://doi.org/10.1002/ cnma.20180​0098

13

126

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



146. Abradelo DG, Cao R, Schlecht S (2013) One-to-one laccase-gold nanoparticle conjugates: molecular recognition and activity enhancement. RSC Adv 3:21461–21465. https​://doi.org/10.1039/c3ra4​ 3192d​ 147. Antuch M, Abradelo DG, Cao R (2014) Bioelectrocatalytic reduction of O ­ 2 at a supramolecularly associated laccase electrode. New J Chem 38:386–390. https​://doi.org/10.1039/c3nj0​1143g​ 148. Johnsen KB, Bak M, Kempen PJ et al (2018) Antibody affinity and valency impact brain uptake of transferrin receptor-targeted gold nanoparticles. Theranostics 8:3416–3436. https​://doi. org/10.7150/thno.25228​ 149. Villegas-Serralta E, Zavala O, Flores-Urquizo IA et al (2018) Detection of HER2 through antibody immobilization is influenced by the properties of the magnetite nanoparticle coating. J Nanomater 2018:1–9. https​://doi.org/10.1155/2018/75716​13 150. Mustafaoglu N, Kiziltepe T, Bilgicer B (2017) Site-specific conjugation of an antibody on a gold nanoparticle surface for one-step diagnosis of prostate specific antigen with dynamic light scattering. Nanoscale 9:8684–8694. https​://doi.org/10.1039/c7nr0​3096g​ 151. Presnova G, Presnov D, Krupenin V et al (2017) Biosensor based on a silicon nanowire field-effect transistor functionalized by gold nanoparticles for the highly sensitive determination of prostate specific antigen. Biosens Bioelectron 88:283–289. https​://doi.org/10.1016/j.bios.2016.08.054 152. Peterson RD, Chen W, Cunningham BT, Andrade JE (2015) Enhanced sandwich immunoassay using antibody-functionalized magnetic iron-oxide nanoparticles for extraction and detection of soluble transferrin receptor on a photonic crystal biosensor. Biosens Bioelectron 74:815–822. https​ ://doi.org/10.1016/j.bios.2015.07.050 153. Denison MIJ, Raman S, Duraisamy N et  al (2015) Preparation, characterization and application of antibody-conjugated magnetic nanoparticles in the purification of begomovirus. RSC Adv 5:99820–99831. https​://doi.org/10.1039/c5ra1​7982c​ 154. Treerattrakoon K, Chanthima W, Apiwat C et al (2017) Oriented conjugation of antibodies against the epithelial cell adhesion molecule on fluorescently doped silica nanoparticles for flow-cytometric determination and in vivo imaging of EpCAM, a biomarker for colorectal cancer. Microchim Acta 184:1941–1950. https​://doi.org/10.1007/s0060​4-017-2211-6 155. Santra S, Kaittanis C, Grimm J, Perez JM (2009) Drug/dye-loaded, multifunctional iron oxide nanoparticles for combined targeted cancer therapy and dual optical/magnetic resonance imaging. Small 5:1862–1868. https​://doi.org/10.1002/smll.20090​0389 156. Jain AK, Thanki K, Jain S (2013) Co-encapsulation of tamoxifen and quercetin in polymeric nanoparticles: implications on oral bioavailability, antitumor efficacy, and drug-induced toxicity. Mol Pharm 10:3459–3474. https​://doi.org/10.1021/mp400​311j 157. Granata G, Stracquadanio S, Leonardi M et al (2018) Essential oils encapsulated in polymer-based nanocapsules as potential candidates for application in food preservation. Food Chem 269:286– 292. https​://doi.org/10.1016/j.foodc​hem.2018.06.140 158. Bravo Cadena M, Preston GM, Van der Hoorn RAL et  al (2018) Species-specific antimicrobial activity of essential oils and enhancement by encapsulation in mesoporous silica nanoparticles. Ind Crops Prod 122:582–590. https​://doi.org/10.1016/j.indcr​op.2018.05.081 159. Liakos IL, Iordache F, Carzino R et al (2018) Cellulose acetate—essential oil nanocapsules with antimicrobial activity for biomedical applications. Coll Surf B Biointerfaces 172:471–479. https​:// doi.org/10.1016/j.colsu​rfb.2018.08.069 160. Hossain F, Follett P, Vu KD et  al (2019) Antifungal activity of combined treatments of active methylcellulose-based films containing encapsulated nanoemulsion of essential oils and γ–irradiation: in  vitro and in  situ evaluations. Cellulose 26:1335–1354. https​://doi.org/10.1007/s1057​ 0-018-2135-2 161. Chu Y, Xu T, Gao CC et  al (2019) Evaluations of physicochemical and biological properties of pullulan-based films incorporated with cinnamon essential oil and Tween 80. Int J Biol Macromol 122:388–394. https​://doi.org/10.1016/j.ijbio​mac.2018.10.194 162. Risaliti L, Kehagia A, Daoultzi E et al (2019) Liposomes loaded with Salvia triloba and Rosmarinus officinalis essential oils: in vitro assessment of antioxidant, antiinflammatory and antibacterial activities. J Drug Deliv Sci Technol 51:493–498. https​://doi.org/10.1016/j.jddst​.2019.03.034 163. Park SJ, Garcia CV, Shin GH, Kim JT (2017) Development of nanostructured lipid carriers for the encapsulation and controlled release of vitamin D3. Food Chem 225:213–219. https​://doi. org/10.1016/j.foodc​hem.2017.01.015

Reprinted from the journal

127

13



Topics in Current Chemistry (2020) 378:12

164. Rivas M, Del Valle LJ, Rodríguez-Rivero AM et al (2018) Loading of antibiotic into biocoated hydroxyapatite nanoparticles: smart antitumor platforms with regulated release. ACS Biomater Sci Eng 4:3234–3245. https​://doi.org/10.1021/acsbi​omate​r ials​.8b003​53 165. Khoshakhlagh K, Koocheki A, Mohebbi M, Allafchian A (2017) Development and characterization of electrosprayed Alyssum homolocarpum seed gum nanoparticles for encapsulation of d-limonene. J Colloid Interface Sci 490:562–575. https​://doi.org/10.1016/j.jcis.2016.11.067 166. Zhang F, Khan MA, Cheng H, Liang L (2019) Co-encapsulation of α-tocopherol and resveratrol within zein nanoparticles: impact on antioxidant activity and stability. J Food Eng 247:9–18. https​://doi.org/10.1016/j.jfood​eng.2018.11.021 167. Abbasi F, Samadi F, Jafari SM et  al (2019) Ultrasound-assisted preparation of flaxseed oil nanoemulsions coated with alginate-whey protein for targeted delivery of omega-3 fatty acids into the lower sections of gastrointestinal tract to enrich broiler meat. Ultrason Sonochem 50:208–217. https​://doi.org/10.1016/j.ultso​nch.2018.09.014 168. Wang S, Chen Y, Wang S et  al (2019) DNA-functionalized metal-organic framework nanoparticles for intracellular delivery of proteins. J Am Chem Soc 141:2215–2219. https​://doi. org/10.1021/jacs.8b127​05 169. Mukai H, Hatanaka K, Yagi N et al (2019) Pharmacokinetic evaluation of liposomal nanoparticle-encapsulated nucleic acid drug: a combined study of dynamic PET imaging and LC/MS/MS analysis. J Control Release 294:185–194. https​://doi.org/10.1016/j.jconr​el.2018.12.006 170. Koyani RD, Vazquez-Duhalt R (2016) Laccase encapsulation in chitosan nanoparticles enhances the protein stability against microbial degradation. Environ Sci Pollut Res 23:18850– 18857. https​://doi.org/10.1007/s1135​6-016-7072-8 171. Ishizuka F, Chapman R, Kuchel RP et  al (2018) Polymeric nanocapsules for enzyme stabilization in organic solvents. Macromolecules 51:438–446. https​://doi.org/10.1021/acs.macro​ mol.7b023​77 172. Hu C, Wu J, Wei T et al (2018) A supramolecular approach for versatile biofunctionalization of magnetic nanoparticles. J Mater Chem B 6:2198–2203. https​://doi.org/10.1039/c8tb0​0490k​ 173. Hirsh SL, Bilek MMM, Nosworthy NJ et  al (2010) A comparison of covalent immobilization and physical adsorption of a cellulase enzyme mixture. Langmuir 26:14380–14388. https​://doi. org/10.1021/la101​9845 174. Jiang B, Dong P, Zheng J (2018) A novel amperometric biosensor based on covalently attached multilayer assemblies of gold nanoparticles, diazo-resins and acetylcholinesterase for the detection of organophosphorus pesticides. Talanta 183:114–121. https​://doi.org/10.1016/j.talan​ ta.2018.02.016 175. Taheran M, Naghdi M, Brar SK et  al (2017) Covalent immobilization of laccase onto nanofibrous membrane for degradation of pharmaceutical residues in water. ACS Sustain Chem Eng 5:10430–10438. https​://doi.org/10.1021/acssu​schem​eng.7b024​65 176. Amin R, Khorshidi A, Shojaei AF et al (2018) Immobilization of laccase on modified ­Fe3O4@ SiO2@Kit-6 magnetite nanoparticles for enhanced delignification of olive pomace bio-waste. Int J Biol Macromol 114:106–113. https​://doi.org/10.1016/j.ijbio​mac.2018.03.086 177. Iriarte-Mesa C, Díaz-Castañón S, Abradelo DG (2019) Facile immobilization of Trametes versicolor laccase on highly monodisperse superparamagnetic iron oxide nanoparticles. Coll Surf B Biointerfaces 181:470–479. https​://doi.org/10.1016/j.colsu​r fb.2019.05.012 178. Zhu X, Li Y, Yang G et  al (2019) Covalent immobilization of alkaline proteinase on aminofunctionalized magnetic nanoparticles and application in soy protein hydrolysis. Biotechnol Prog 35:e2756. https​://doi.org/10.1002/btpr.2756 179. Zhu YT, Ren XY, Liu YM et  al (2014) Covalent immobilization of porcine pancreatic lipase on carboxyl-activated magnetic nanoparticles: characterization and application for enzymatic inhibition assays. Mater Sci Eng, C 38:278–285. https​://doi.org/10.1016/j.msec.2014.02.011 180. Chiou SH, Wu WT (2004) Immobilization of Candida rugosa lipase on chitosan with activation of the hydroxyl groups. Biomaterials 25:197–204. https​://doi.org/10.1016/S0142​ -9612(03)00482​-4 181. Barbosa M, Vale N, Costa FMTA et al (2017) Tethering antimicrobial peptides onto chitosan: optimization of azide-alkyne “click” reaction conditions. Carbohydr Polym 165:384–393. https​://doi. org/10.1016/j.carbp​ol.2017.02.050 182. Su Z, Xu H, Xu X et al (2017) Effective covalent immobilization of quinone and aptamer onto a gold electrode via thiol addition for sensitive and selective protein biosensing. Talanta 164:244– 248. https​://doi.org/10.1016/j.talan​ta.2016.11.049

13

128

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12



183. Liu C, Saeki D, Matsuyama H (2017) A novel strategy to immobilize enzymes on microporous membranes: via dicarboxylic acid halides. RSC Adv 7:48199–48207. https​://doi.org/10.1039/c7ra1​ 0012d​ 184. Wu X, Wei P, Zhu X et  al (2017) Effect of immobilization on the antimicrobial activity of a cysteine-terminated antimicrobial Peptide Cecropin P1 tethered to silica nanoparticle against E. coli O157 : H7 EDL933. Coll Surf B: Biointerfaces 156:305–312 185. Min K, Yun G, Jang Y et al (2016) Communications nanocapsules hot paper covalent self-assembly and one-step photocrosslinking of tyrosine- rich oligopeptides to form diverse nanostructures. Angewandte. 742:6925–6928. https​://doi.org/10.1002/anie.20160​1675 186. Nogueira F, Karumidze N, Kusradze I et  al (2017) Immobilization of bacteriophage in wounddressing nanostructure. Nanomed Nanotechnol Biol Med 13:2475–2484. https​://doi.org/10.1016/j. nano.2017.08.008 187. Costa FMTA, Maia SR, Gomes PAC, Martins MCL (2015) Biomaterials Dhvar5 antimicrobial peptide (AMP) chemoselective covalent immobilization results on higher antiadherence effect than simple physical adsorption. Biomaterials 52:531–538. https​://doi.org/10.1016/j.bioma​teria​ ls.2015.02.049 188. Borzooeian Z, Taslim ME, Borzooeian G et  al (2017) Activity and stability analysis of covalent conjugated lysozyme-single walled carbon nanotubes: potential biomedical and industrial applications. RSC Adv 7:48692–48701. https​://doi.org/10.1039/C7RA0​7189B​ 189. Tang A, Zhang Y, Wei T et  al (2018) Immobilization of candida cylindracea lipase by covalent attachment on glu-modified bentonite. Appl Biochem Biotechnol 187:870–883. https​://doi. org/10.1007/s1201​0-018-2838-8 190. Xia N, Xing Y, Wang G et  al (2013) Probing of EDC/NHSS-mediated covalent coupling reaction by the immobilization of electrochemically active biomolecules. Int J Electrochem Sci 8:2459–2467 191. Kamra T, Chaudhary S, Xu C et al (2016) Covalent immobilization of molecularly imprinted polymer nanoparticles on a gold surface using carbodiimide coupling for chemical sensing. J Coll Interface Sci 461:1–8. https​://doi.org/10.1016/j.jcis.2015.09.009 192. Kazenwadel F, Wagner H, Rapp BE, Franzreb M (2015) Optimization of enzyme immobilization on magnetic microparticles using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) as a crosslinking agent. Anal Methods 7:10291–10298. https​://doi.org/10.1039/c5ay0​2670a​ 193. Kühn C, von Oesen T, Hanschen FS, Rohn S (2018) Determination of isothiocyanate-protein conjugates in milk and curd after adding garden cress (Lepidium sativum L.). Food Res Int 108:621– 627. https​://doi.org/10.1016/j.foodr​es.2018.04.001 194. Raj J, Herzog G, Manning M et  al (2009) Surface immobilisation of antibody on cyclic olefin copolymer for sandwich immunoassay. Biosens Bioelectron 24:2654–2658. https​://doi. org/10.1016/j.bios.2009.01.026 195. Chaiyasan W, Praputbut S, Kompella UB et al (2017) Penetration of mucoadhesive chitosan-dextran sulfate nanoparticles into the porcine cornea. Coll Surf B Biointerfaces 149:288–296. https​:// doi.org/10.1016/j.colsu​rfb.2016.10.032 196. Jastrzębska A, Piasta A, Krzemiński M, Szłyk E (2018) Application of 3,5-bis-(trifluoromethyl) phenyl isothiocyanate for the determination of selected biogenic amines by LC-tandem mass spectrometry and 19F NMR. Food Chem 239:225–233. https​://doi.org/10.1016/j.foodc​ hem.2017.06.100 197. Cui Q, Hou Y, Hou J et al (2013) Preparation of functionalized alkynyl magnetic microspheres for the selective enrichment of cell glycoproteins based on click chemistry. Biomacromol 14:124–131. https​://doi.org/10.1021/bm301​477z 198. Gruttner C, Muller K, Teller J (2013) Comparison of strain-promoted alkyne-azide cycloaddition with established methods for conjugation of biomolecules to magnetic nanoparticles. IEEE Trans Magn 49:172–176. https​://doi.org/10.1109/TMAG.2012.22185​81 199. Meloni MM, Barton S, Chivu A et  al (2018) Facile, productive, and cost-effective synthesis of a novel tetrazine-based iron oxide nanoparticle for targeted image contrast agents and nanomedicines. J Nanoparticle Res 20:265. https​://doi.org/10.1007/s1105​1-018-4370-8 200. Lin PC, Ueng SH, Yu SC et  al (2007) Surface modification of magnetic nanoparticle via Cu(I)catalyzed alkyne-azide [2 + 3] cycloaddition. Org Lett 9:2131–2134. https​://doi.org/10.1021/ol070​ 588f 201. Finetti C, Sola L, Pezzullo M et al (2016) Click chemistry immobilization of antibodies on polymer coated gold nanoparticles. Langmuir 32:7435–7441. https​://doi.org/10.1021/acs.langm​uir.6b011​42 Reprinted from the journal

129

13



Topics in Current Chemistry (2020) 378:12

202. Luo X, Morrin A, Killard AJ, Smyth MR (2006) Application of nanoparticles in electrochemical sensors and biosensors. Electroanalysis 18:319–326. https​://doi.org/10.1002/elan.20050​ 3415 203. Diao P, Guo M, Zhang Q (2008) How does the particle density affect the electrochemical behavior of gold nanoparticle assembly? J Phys Chem C 112:7036–7046. https​://doi.org/10.1021/ jp077​653n 204. Young SL, Kellon JE, Hutchison JE (2016) Small gold nanoparticles interfaced to electrodes through molecular linkers: a platform to enhance electron transfer and increase electrochemically active surface area. J Am Chem Soc 138:13975–13984. https​://doi.org/10.1021/jacs.6b076​74 205. Bartlett PN, Al-Lolage FA (2018) There is no evidence to support literature claims of direct electron transfer (DET) for native glucose oxidase (GOx) at carbon nanotubes or graphene. J Electroanal Chem 819:26–37. https​://doi.org/10.1016/j.jelec​hem.2017.06.021 206. Huerta-Miranda GA, Arrocha-Arcos AA, Miranda-Hernández M (2018) Gold nanoparticles/4aminothiophenol interfaces for direct electron transfer of horseradish peroxidase: enzymatic orientation and modulation of sensitivity towards hydrogen peroxide detection. Bioelectrochemistry 122:77–83. https​://doi.org/10.1016/j.bioel​echem​.2018.03.004 207. Sanaeifar N, Rabiee M, Abdolrahim M et al (2017) A novel electrochemical biosensor based on ­Fe3O4 nanoparticles-polyvinyl alcohol composite for sensitive detection of glucose. Anal Biochem 519:19–26. https​://doi.org/10.1016/j.ab.2016.12.006 208. Márquez A, Jiménez-Jorquera C, Domínguez C, Muñoz-Berbel X (2017) Electrodepositable alginate membranes for enzymatic sensors: an amperometric glucose biosensor for whole blood analysis. Biosens Bioelectron 97:136–142. https​://doi.org/10.1016/j.bios.2017.05.051 209. Moses Phiri M, Wingrove Mulder D, Mason S, Christiaan Vorster B (2019) Facile immobilization of glucose oxidase onto gold nanostars with enhanced binding affinity and optimal function. R Soc Open Sci 6:190205. https​://doi.org/10.1098/rsos.19020​5 210. Aggarwal V, Malik J, Prashant A et al (2016) Amperometric determination of serum total cholesterol with nanoparticles of cholesterol esterase and cholesterol oxidase. Anal Biochem 500:6–11. https​://doi.org/10.1016/j.ab.2016.01.019 211. Pakapongpan S, Poo-arporn RP (2017) Self-assembly of glucose oxidase on reduced graphene oxide-magnetic nanoparticles nanocomposite-based direct electrochemistry for reagentless glucose biosensor. Mater Sci Eng, C 76:398–405. https​://doi.org/10.1016/j.msec.2017.03.031 212. Rassas I, Braiek M, Bonhomme A et  al (2019) Highly sensitive voltammetric glucose biosensor based on glucose oxidase encapsulated in a chitosan/kappa-carrageenan/gold nanoparticle bionanocomposite. Sensors 19:154. https​://doi.org/10.3390/s1901​0154 213. Kawde AN, Aziz MA, El-Zohri M et  al (2017) Cathodized gold nanoparticle-modified graphite pencil electrode for non-enzymatic sensitive voltammetric detection of glucose. Electroanalysis 29:1214–1221. https​://doi.org/10.1002/elan.20160​0709 214. Baig N, Sajid M, Saleh TA (2019) Recent trends in nanomaterial-modified electrodes for electroanalytical applications. TrAC Trends Anal Chem 111:47–61. https​://doi.org/10.1016/j. trac.2018.11.044 215. Zou Y, Liang J, She Z, Kraatz HB (2019) Gold nanoparticles-based multifunctional nanoconjugates for highly sensitive and enzyme-free detection of E. coli K12. Talanta 193:15–22. https​://doi. org/10.1016/j.talan​ta.2018.09.068 216. Hoo XF, Abdul Razak K, Ridhuan NS et  al (2019) Electrochemical glucose biosensor based on ZnO nanorods modified with gold nanoparticles. J Mater Sci Mater Electron 30:7460–7470. https​ ://doi.org/10.1007/s1085​4-019-01059​-9 217. Buk V, Pemble ME, Twomey K (2018) Fabrication and evaluation of a carbon quantum dot/gold nanoparticle nanohybrid material integrated onto planar micro gold electrodes for potential bioelectrochemical sensing applications. Electrochim Acta 293:307–317. https​://doi.org/10.1016/j.elect​ acta.2018.10.038 218. Buk V, Pemble ME (2019) A highly sensitive glucose biosensor based on a micro disk array electrode design modified with carbon quantum dots and gold nanoparticles. Electrochim Acta 298:97– 105. https​://doi.org/10.1016/j.elect​acta.2018.12.068 219. Chen Z, Lai G, Liu S, Yu A (2018) Ultrasensitive electrochemical aptasensing of kanamycin antibiotic by enzymatic signal amplification with a horseradish peroxidase-functionalized gold nanoprobe. Sensors Actuators, B Chem 273:1762–1767. https​://doi.org/10.1016/j.snb.2018.07.102 220. Chiodo F, Marradi M (2015) Gold nanoparticles as carriers for synthetic glycoconjugate vaccines. Methods Mol Biol 1331:159–171. https​://doi.org/10.1007/978-1-4939-2874-3_10

13

130

Reprinted from the journal

Topics in Current Chemistry (2020) 378:12 221. Zhang X, Huang C, Jiang Y et al (2016) An electrochemical glycan biosensor based on a thioninebridged multiwalled carbon nanotube/gold nanoparticle composite-modified electrode. RSC Adv 6:112981–112987. https​://doi.org/10.1039/c6ra2​3710j​ 222. Eivazzadeh-Keihan R, Pashazadeh-Panahi P, Baradaran B et al (2018) Recent advances on nanomaterial based electrochemical and optical aptasensors for detection of cancer biomarkers. TrAC Trends Anal Chem 100:103–115. https​://doi.org/10.1016/j.trac.2017.12.019 223. Nguyen NV, Jen CP (2019) Selective detection of human lung adenocarcinoma cells based on the aptamer-conjugated self-assembled monolayer of gold nanoparticles. Micromachines 10:195. https​ ://doi.org/10.3390/mi100​30195​ 224. Lima D, Inaba J, Clarindo Lopes L et al (2019) Label-free impedimetric immunosensor based on arginine-functionalized gold nanoparticles for detection of DHEAS, a biomarker of pediatric adrenocortical carcinoma. Biosens Bioelectron 133:86–93. https​://doi.org/10.1016/j.bios.2019.02.063 225. Maduraiveeran G, Sasidharan M, Ganesan V (2018) Electrochemical sensor and biosensor platforms based on advanced nanomaterials for biological and biomedical applications. Biosens Bioelectron 103:113–129. https​://doi.org/10.1016/j.bios.2017.12.031 226. Abbaspour A, Norouz-Sarvestani F, Noori A, Soltani N (2015) Aptamer-conjugated silver nanoparticles for electrochemical dual-aptamer-based sandwich detection of staphylococcus aureus. Biosens Bioelectron 68:149–155. https​://doi.org/10.1016/j.bios.2014.12.040 227. Ashrafi H, Hassanpour S, Saadati A et  al (2019) Sensitive detection and determination of benzodiazepines using silver nanoparticles-N-GQDs ink modified electrode: a new platform for modern pharmaceutical analysis. Microchem J 145:1050–1057. https​://doi.org/10.1016/j.micro​ c.2018.12.017 228. Roushani M, Shahdost-fard F (2015) A novel ultrasensitive aptasensor based on silver nanoparticles measured via enhanced voltammetric response of electrochemical reduction of riboflavin as redox probe for cocaine detection. Sensors Actuators B Chem 207:764–771. https​://doi. org/10.1016/j.snb.2014.10.131 229. Aftab S, Kurbanoglu S, Ozcelikay G et  al (2019) Carbon quantum dots co-catalyzed with multiwalled carbon nanotubes and silver nanoparticles modified nanosensor for the electrochemical assay of anti-HIV drug Rilpivirine. Sensors Actuators, B Chem 285:571–583. https​://doi. org/10.1016/j.snb.2019.01.094 230. Meng F, Sun H, Huang Y et al (2019) Peptide cleavage-based electrochemical biosensor coupling graphene oxide and silver nanoparticles. Anal Chim Acta 1047:45–51. https​://doi.org/10.1016/j. aca.2018.09.053 231. Elhakim HKA, Azab SM, Fekry AM (2018) A novel simple biosensor containing silver nanoparticles/propolis (bee glue) for microRNA let-7a determination. Mater Sci Eng, C 92:489–495. https​:// doi.org/10.1016/j.msec.2018.06.063 232. Chen S, Qamar AZ, Asefifeyzabadi N et al (2019) Hand-fabricated CNT/AgNPs electrodes using wax-on-plastic platforms for electro-immunosensing application. Sci Rep 9:6131. https​://doi. org/10.1038/s4159​8-019-42644​-6 233. Katz E, Willner I (2005) Switching of directions of bioelectrocatalytic currents and photocurrents at electrode surfaces by using hydrophobic magnetic nanoparticles. Angew Chemie Int Ed 44:4791–4794. https​://doi.org/10.1002/anie.20050​1126 234. Yang L, Ren X, Tang F, Zhang L (2009) A practical glucose biosensor based on ­Fe3O4 nanoparticles and chitosan/nafion composite film. Biosens Bioelectron 25:889–895. https​://doi. org/10.1016/j.bios.2009.09.002 235. Antuch M, Matos-Peralta Y, Llanes D et al (2019) Bimetallic ­Co2+ and ­Mn2+ hexacyanoferrate for hydrogen peroxide electrooxidation and its application in a highly sensitive cholesterol biosensor. ChemElectroChem 6:1567–1573. https​://doi.org/10.1002/celc.20190​0190 236. Tian L, Qi J, Qian K et al (2018) An ultrasensitive electrochemical cytosensor based on the magnetic field assisted binanozymes synergistic catalysis of ­Fe3O4 nanozyme and reduced graphene oxide/molybdenum disulfide nanozyme. Sensors Actuators B Chem 260:676–684. https​://doi. org/10.1016/j.snb.2018.01.092 237. Teymourian H, Salimi A, Khezrian S (2017) Development of a new label-free, indicator-free strategy toward ultrasensitive electrochemical DNA Biosensing Based on ­Fe3O4 nanoparticles/reduced graphene oxide composite. Electroanalysis 29:409–414. https​://doi.org/10.1002/elan.20160​0336 238. Tufa LT, Oh S, Tran VT et al (2018) Electrochemical immunosensor using nanotriplex of graphene quantum dots, ­Fe3O4 and Ag nanoparticles for tuberculosis. Electrochim Acta 290:369–377. https​ ://doi.org/10.1016/j.elect​acta.2018.09.108 Reprinted from the journal

131

13



Topics in Current Chemistry (2020) 378:12

239. Shamsipur M, Emami M, Farzin L, Saber R (2018) A sandwich-type electrochemical immunosensor based on in situ silver deposition for determination of serum level of HER2 in breast cancer patients. Biosens Bioelectron 103:54–61. https​://doi.org/10.1016/j.bios.2017.12.022 240. Kumar S, Umar M, Saifi A et  al (2019) Electrochemical paper based cancer biosensor using iron oxide nanoparticles decorated PEDOT: PSS. Anal Chim Acta 1056:135–145. https​://doi. org/10.1016/j.aca.2018.12.053 241. Chikhaliwala P, Rai R, Chandra S (2019) Simultaneous voltammetric immunodetection of alphafetoprotein and glypican-3 using a glassy carbon electrode modified with magnetite-conjugated dendrimers. Microchim Acta 186:255. https​://doi.org/10.1007/s0060​4-019-3354-4

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

13

132

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35 https://doi.org/10.1007/s41061-020-0296-6 REVIEW

Quantum Dot Bioconjugates for Diagnostic Applications María Díaz‑González1 · Alfredo de la Escosura‑Muñiz1 · Maria Teresa Fernandez‑Argüelles2 · Francisco Javier García Alonso3 · Jose Manuel Costa‑Fernandez2  Received: 30 December 2019 / Accepted: 29 February 2020 / Published online: 26 March 2020 © Springer Nature Switzerland AG 2020

Abstract Quantum dots (QDs) are a special type of engineered nanomaterials with outstanding optoelectronic properties that make them as a very promising alternative to conventional luminescent dyes in biomedical applications, including biomolecule (BM) targeting, luminescence imaging and drug delivery. A key parameter to ensure successful biomedical applications of QDs is the appropriate surface modification, i.e. the surface of the nanomaterials should be modified with the appropriate functional groups to ensure stability in aqueous solutions and it should be conjugated with recognition elements capable of ensuring an efficient tagging of the BMs of interest. In this review we summarize the most relevant strategies used for surface modification of QDs and for their conjugation to BMs in preparation of their application in nanoplatforms for luminescent BM sensing and imaging-guided targeting. The applications of conjugations of photoluminescent QDs with different BMs in both in vitro and in vivo chemical sensing, immunoassays or luminescence imaging are reviewed. Recent progress in the application of functionalized QDs in ultrasensitive detection in bioanalysis, diagnostics and imaging strategies are reported. Finally, some key future research goals in the progress of bioconjugation of QDs for diagnosis are identified, including novel synthetic approaches, the need for exhaustive characterization of bioconjugates and the design of signal amplification schemes. Keywords  Biosensing · Diagnostics · Luminescence · Nanoparticles · Quantum dots · Surface functionalization

Chapter 5 was originally published as Díaz‑González, M., Escosura‑Muñiz, A. de la, Fernandez‑Argüelles, M. T., Alonso, F. J. G. & Costa‑Fernandez, J. M. Topics in Current Chemistry (2020) 378: 35. https://doi.org/10.1007/s41061-020-0296-6. * Jose Manuel Costa‑Fernandez [email protected] Extended author information available on the last page of the article Reprinted from the journal

133

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:35

1 Introduction Nanotechnology encompasses the fabrication, characterization, manipulation and application of materials that have at least one dimension within the range 1–100  nm. When the size of the material is below this threshold, the material behaves differently from the same material with macroscopic dimensions due to the quantum confinement of the electrons when the dimensions are smaller than the Bohr radius, giving rise to unique and extraordinary physicochemical properties [1]. Nanomaterials can be classified according to different criteria, with the most frequent criterion based on the dimensions of the nanomaterial. Three-dimensional (3-D) nanomaterials are those with three dimensions larger than the nanometer scale, but which are composed of individual building blocks that are within the nanometer scale range, such as nanocomposites. Two-dimensional (2-D) nanomaterials are those that present one dimension in the nanometer scale and include, for example, thin films or nanocoatings. One-dimensional (1-D) nanomaterials possess two dimensions within the range of 1–100  nm and only one dimension larger than the nanometer scale; these nanomaterials include, among others, nanotubes, nanorods and nanowires. Finally, zero-dimensional (0-D) nanomaterials are those with three dimensions within the range 1–100 nm, including nanoparticles (NPs) such as metal NPs, semiconductor quantum dots (QDs), and carbon-based QDs (CQDs) [2]. The development of novel nanomaterials has gained increasing interest in recent decades due to their fascinating physicochemical properties, which have a a great potential for application in different research areas and industries, such as (bio)analytical chemistry [3, 4], water treatment systems [5], catalysis, electrocatalysis [6–9], cancer treatment [10], energy storage devices [11], among others. Although little information is currently available on the production of QDs, it was possible to estimate worldwide and Europe-wide production and use of ten different nanomaterials, including QDs, from a 2012 survey sent to companies producing and using engineered nanomaterials, with the results indicating that the estimated production of QDs was about 10 t/year or lower [12]. Inorganic semiconductor nanocrystals, or QDs, have demonstrated a range of unique optoelectronic properties and represent novel, attractive options in many biomedical applications [13–16]. For example, QDs have been widely employed for fluorescence sensing and bioimaging due to their exceptional photoluminescent characteristics, including the capability to tune the emission wavelength just by controlling the size of the NP. Although conventional organic fluorescent molecules are widely used in bioimaging applications, QDs are superior luminescence tags in terms of their photophysical properties, namely, QDs have broad excitation spectra and narrow and sharp emission spectra and large Stokes shifts (> 100  nm). Such optoelectronic properties are of great value in multiplexed applications [17, 18] as by using a single light source it is possible to simultaneously excite multiple QDs of different sizes (multiple emission peaks). Additionally, QDs have very high molar adsorption coefficients, as well as higher

13

134

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35

quantum yields than organic fluorophores. Consequently, fluorescent NPs are nearly 20-fold brighter and many thousand-fold more stable against photobleaching than conventional organic dyes [19, 20]. Such exceptional optical behavior justifies the rapid emergence of QDs as valuable photoluminescent probes in many analytical applications. In particular, the use of NPs for diagnostic purposes is increasing exponentially due to their highly valuable optoelectronic properties and small size. Additionally, recent advances in such areas as surface modification and functionalization have given rise to the improved colloidal stability of NPs in complex media and biological buffers and biocompatibility, while allowing the NPs to bond to recognition elements [21]. In this context, especially relevant are the developments in the synthesis of NPs with interesting optical properties that overcome the limitations of traditional organic dyes, including the synthesis of gold NPs (AuNPs), semiconductor QDs, gold nanoclusters, silver nanoclusters, rare-earth-based NPs, carbon dots and dyeloaded NPs [22]. However, despite the exceptional properties demonstrated by QDs developed to date, their applications in clinical analysis, especially during the first decades of development, can still be considered to be somehow limited, in part due to their low targeting efficiency and eventual high toxicity, both of which could hinder their application in in vivo imaging. These limitations underlie the many research efforts to develop QDs exhibiting low biological toxicity (e.g. those based on an A ­ g2S core) [23]. Additionally, in the absence of any molecular moiety being attached to the NP surface, QDs generally show nonselective distribution acros different organisms, thus failing to satisfy the minimum requirements for appropriate molecular imaging. Clearly, the development of QD-based nanoprobes requires a previous surface functionalization of the NPs to facilitate the various approaches used in targeting-guided imaging techniques. In addition, to ensure the required biocompatibility of the nanoprobes to be used in in vivo imaging or sensing, an appropriate surface functionalization of the NPs is required. The aim of this review is to highlight advances in the use of QDs in diagnostic applications. In the following sections, we first introduce and briefly describe the main types of QDs used in bioanalysis. This is followed by a section that focuses on the strategies of solubilization and stabilization of QDs in aqueous solutions under physiological conditions and then by a section in which approaches used for the functionalization of QDs with biological molecules are summarized. The functionalization of QDs is a key aspect of their use and a requirement before they can be employed for the detection of analytes in biological matrices. Thereafter, we describe some of the most relevant applications of QD bioconjugates in the optical imaging of biomarkers, including a review of the in  vitro applications of QDs in medicine in which different detection schemes based on the bioconjugation of QDs to antibodies, aptamers, peptides or other types of recognition elements. In this context, we provide an overview of recent advances in the development of lowcost, portable and easy-to-use QD-based biosensing devices for clinical applications (point-of-care). The final section discusses future prospects with the intention to indicate the direction that research on the use of QDs in diagnostic applications is heading. Reprinted from the journal

135

13



Topics in Current Chemistry (2020) 378:35

2 QDs: Nature and Types Despite the many problems encountered in the initial attempts to synthesize colloidal fluorescent semiconductor NPs, such as lack of reproducibility and reduced optical quality, important advances have been achieved in this field. In pioneer studies on the routine preparation of colloidal QDs, the core of the QDs was usually capped with an organic layer that coordinates with core-metal sites and stabilizes the QD surface, thereby preventing an irreversible flocculation (aggregation) of the nanocrystals [24]. Unfortunately, these protective ligands are also hydrophobic, and thus nanocrystals capped with such coatings are not compatible with bioanalytical assay conditions. Consequently, the QD surface should be further modified by attaching the appropriate hydrophilic functional groups to allow dispersion of the QDs in aqueous solutions while maintaining their high photoluminescence quantum yield. Although fluorescence emission is the most exploited property of QDs, doping the core of the nanocrystals with transition elements has been a strategy adopted to provide the QDs with new improved multimodal characteristics for biomedical applications. Additionally, during recent years, many research groups have tried to overcome the problem of eventual cytotoxicity of the more conventional heavymetal based QDs. Approaches based on carbon-based nanomaterials are one of the most promising strategies. In this section, we briefly review the nature and characteristics of QDs typically used in clinical and biomedical applications. 2.1 Semiconductor QDs Quantum dots are spherical semiconductor photoluminescent NPs with a diameter ranging between 2 and 10 nm. Since the dimensions of the NP are smaller than the Bohr radius, the energy levels are quantized. As a consequence, the optoelectronic properties of QDs depend on their size due to quantum confinement effects and differ from the properties observed for the same bulk material [25]. In fact, due to the quantized energy levels, QDs generate an intense emission of photoluminescence: when the semiconductor QD is irradiated with a light source, the absorption causes an electron to move from the lower energy valence band to a higher energy conduction band, following which an electron–hole pair is generated, and its recombination gives rise to the emission of intense photoluminescence. Other optical features that make QDs very appealing for use in analytical applications include broad absorption spectra, narrow and symmetric emission bands that can be tuned by changing the composition and size of the NP (see Fig. 1), large Stoke shifts and high photostability [26, 27]. The energy band gap decreases with increases in QD diameter, and as a result the emission wavelength shifts to longer wavelengths. Hence, for QDs with the same composition, the emission can be tuned by just modifying the size of the NP (referred to as size-dependent emission). In addition, the QD can be synthesized with different semiconductors, such as CdS, CdSe, CdTe, ZnS, ZnSe,

13

136

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



Normalized Fluorescence Intensity

1

0.8

0.6

0.4

0.2

0 480

500

520

540

560

580

600

620

640

wavelength (nm)

Fig. 1  Size-tuneable fluorescence spectra of CdSe quantum dots (QDs) of different diameter sizes. Bottom bar shows images of the colloidal suspensions of the differently sized QDs under UV light. Reprinted from Fernandez-Argüelles et al. [28], copyright 2010, with permission from Wiley

PbS, InP, among others, all of which also affect their spectral properties. Such features of QDs are very attractive because emission in a wide range of spectra, from the ultraviolet (UV) up to the near-infrared (NIR), can be obtained by simpley changing the composition and size of the QD. Additionally, all QDs generate broad absorption bands, meaning that a single excitation source can be used to efficiently excite QDs with different emission wavelengths, a characteristic which, in addition to their high luminescent quantum yields and long photostability (e.g.: CdS/ZnS QDs are nearly 20-fold brighter and 100-fold more stable than the widely used Rhodamine 6G [19]), is of great interest in multianalyte detection [16, 29]. The synthesis of semiconductor QDs can be performed in both organic hydrophobic solvent and aqueous media, although the synthetic routes that generate QDs with the best optoelectronic properties are those carried out in nonpolar solvents and using hydrophobic ligands. As a result, while the QDs obtained through these routes present outstanding photoluminescent properties, they tend to aggregate and precipitate in aqueous solutions. This renders it necessary to modify their surface with molecules that present hydrophilic groups oriented towards the medium, a configuration which bestows the QDs with good colloidal stability in aqueous media. The most common strategies to transfer QDs from hydrophobic organic media to aqueous solution are summarized in section Stabilization Strategies for QDs in Aqueous Media. Reprinted from the journal

137

13



Topics in Current Chemistry (2020) 378:35

2.2 Metal‑Doped QDs In recent years, advances made in the rational design of nanomaterials have contributed to the development of hybrid NPs that combine the interesting size- and shape-dependent properties of semiconductor QDs with a long-lived phosphorescence-type emission. In this context, the incorporation of suitable atoms or ions into host lattices has generated a novel type of QDs that are very promising for use in bioanalytical applications [30]. ZnS, ZnO, ZnSe, CdS and CdSe QDs can be used as host lattices into which other transition-metal and lanthanide ions, including ­Mn2+, ­Cu2+, ­Co2+, ­Ni2+, ­Ag+, ­Pb2+, ­Cr3+, ­Eu3+, ­Tb3+, ­Sm3+ and ­Er3+, are incorporated as dopant agents, giving rise to luminescent QDs with novel properties [31]. Consequently, host lattices such as as ZnS and ZnO are being widely studied because, first, they do not contain toxic metals and thus potentially have a lower toxicity and, second, they are characterized by a larger energy band gap, which allows the incorporation of more doping agents, which is very attractive in terms of developing dual-doped QDs [32, 33]. The introduction of the dopant typically increases the photoluminescence lifetime of the QD (see Fig. 2), producing a phosphorescence-like emission that overcomes the limitations of fluorescent NPs or dyes due to the removal of the fluorescence background commonly found in biosensing and bioimaging applications [14]. Most of the synthetic routes described to incorporate the doping agent into the host structure are based on wet chemistry procedures carried out in organic media, typically under high temperatures, or in aqueous media, through precipitation or microemulsion methods in order to control the size and shape of the nanocrystal as well as obtain a homogeneous distribution of the dopant in the host matrix.

Fig. 2  a Excitation and emission spectra of colloidal Mn:ZnS QDs (solid line) and of colloidal ZnS QDs (broken line), b decay curve of luminescence emission of colloidal Mn:ZnS QDs with a first lifetime component in the range of 0.3 ms and a second longer lifetime component of around 2.1 ms Reprinted from [34], copyright 2012, with permission from Elsevier

13

138

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



2.3 Carbon‑Based QDs (C‑dots) Inorganic semiconductor QDs have been intensely evaluated as luminescent nanomaterials for bioanalytical applications, but carbon-based NPs, such as CQDs and graphene QDs (GQDs), have also drawn attention as attractive alternatives due to their high photoluminescence quantum yields, low photobleaching effects, high biocompatibility and low toxicity, while avoiding the use of heavy metals commonly found in semiconductor QDs [35]. Additionally, carbon-based NPs possess an exceptional colloidal stability in aqueous media as a consequence of their small size, since Brownian motion provides sufficient energy to inhibit aggregation between them [36, 37]. The most remarkable property of CQDs is most likely their excitation wavelength-dependent fluorescence emission (see Fig.  3), which makes them excellent alternative NPs for optical imaging applications [38]. The principle of such characteristic emission is not fully understood, and the origin of the fluorescence emission of CQDs remains a topic of heated discussion. In this context, CQDs obtained through different synthetic routes or using different precursors present different optical behaviors, suggesting that CQDs are quite complex. In fact, the scientific community has yet to agree on an explanation of the optical properties of CQDs, which have been variously attributed to surface state emission, intrinsic band emission, triple ground state emission, dipole emission involving electron–phonon coupling, transition from surface electrons to valence holes, selftrapped excitons and even to the presence of small organic molecules [39, 40]. The synthesis of CQDs has typically employed graphite as the carbon source, and a surface passivation of the CQDs is frequently necessary in order to obtain better fluorescence properties. However, green methods for the synthesis of CQDs based on the use of natural precursors are gaining in importance, as in addition to these synthesis methods being cost effective and environmentally friendly, the CQDs obtained do not require any surface modification, present high photoluminescence and have excellent stability in aqueous media [41].

(a)

8.0x106

Fluorescence Intensity

Fluorescence Intensity

8.0x106

290 nm 300 nm 310 nm 320 nm 330 nm 340 nm 350 nm 360 nm 370 nm 380 nm 390 nm

6.0x106

4.0x106

2.0x106

0.0

400

500

600

Wavelength (nm)

390 nm 400 nm 410 nm 420 nm 430 nm 440 nm 450 nm 460 nm 470 nm 480 nm 490 nm 500 nm 510 nm 520 nm 530 nm 540 nm 550 nm 560 nm 570 nm 580 nm 590 nm 600 nm

6.0x106

4.0x106

2.0x106

0.0 400

700

(b)

500

600

700

Wavelength (nm)

Fig. 3  Emission spectra of black pepper carbon QDs (CQDs) under different excitation wavelengths. a excitation wavelength from 290 to 390 nm, b excitation wavelength from 390 to 600 nm. Reprinted with permission from Vasimalai et al. [39], copyright 2018 Beilstein-Institut Reprinted from the journal

139

13



Topics in Current Chemistry (2020) 378:35

The potential of CQDs for use in diagnostic applications has increased recently with the metal doping of NPs with N and lanthanides (e.g. Gd and Yb), resulting in co-doped nanomaterials exhibiting not only strong fluorescence but also high constrast capabilities in magnetic resonance imaging (MRI) and computed tomography (CT) [42]. This simple approach has enabled the design of multimodal QDs for bioimaging applications. However, studies on the surface modification of these co-doped NPs with appropriate recognition biomolecules (BMs) for targeted bioimaging are still needed.

3 Stabilization Strategies for QDs in Aqueous Media As mentioned in a previous section, conventional high-quality fluorescent QDs are commonly synthesized in organic solvents at high temperature. However, the QDs must be made water-compatible (stable in aqueous and biological media so that they maintain their optoelectronic properties) if the intenstion is to use them in bioanalytical applications. To this end, surface modification of the QDs after synthesis is a must; as well, QDs should have functional groups available on their surface for further bioconjugation to BMs. The appropriate QD surface passivation also can solve some of the problems typically affecting these NPs. First, crystalline NPs can easily form surface defects that quench the fluorescence properties of naked QDs [43]. Second, naked QDs can suffer from surface oxidation, photochemical degradation and/or the leaching of metal ions from the NP core after exposure to ionic or biological media, which affects their optoelectronic properties and produces undesirable cytotoxicity [44]. Thus, modification of the QD surface with the appropriate ligands is essential not only to stabilize the NPs in physiological media (particularly important if they are going to be used in clinical applications) but also to reduce nanocrystal surface defects, thereby minimizing QD reactivity and toxicity. Moreover, despite the significant progress achieved in the synthesis of QDs, biological uses of QDs require that such NPs be modified into biocompatible probes. In this context, the availability of robust and versatile NP surface chemistries are invaluable strategies to achieve stabilization of the QDs in biological buffers while preserving their original photophysical properties and providing adequate reactive groups for further bioconjugations. The three main strategies employed for hydrophilization of QDs (based on the attachment of polar functional groups to the surface of the QD) are summarized in Fig. 4. As shown in Fig. 4, a universal and simple approach is based on ligand exchange of the original hydrophobic coating of the QDs (e.g. trioctylphosphine oxide [TOPO] chains). In this method, the original coating is removed and replaced with bifunctional molecules that often bind to the QD surface (e.g. through a thiol end) and which have a hydrophilic functional group on the other end (such as carboxyl or sulfonic acids) that provides the required NP solubility in aqueous and polar media and is also available for further bioconjugation [19]. Bidentate ligands, such as dithiothreitol (DTT) or dihydrolipoic acid (DHLA), as well as oligomeric phosphines, peptides and crosslinked dendrons are widely used to obtain QDs that are stable in an aqueous environment [45].

13

140

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



Fig. 4  The three QD phase-transfer approaches commonly used for aqueous stabilization of bare nanoparticles (NPs): the ligand or cap exchange process; bonding of amphiphilic polymers and phospholipids to hydrophobic groups on the surface of the QDs; and surface silanization of the core NP. Reprinted from Karakoti et al. [46], copyright 2015, with permission from Elsevier

Another method commonly used to achieve QD stability in an aqueous environment is based on the well-known silica chemistry used for inorganic encapsulation of the hydrophilic structures (e.g. surface silanization) of the QDs through the generation of a silica shell around the NP surface [46]. This is a very attractive approach to achieve water stabilization of QDs as the silica surface is non-toxic, chemically inert and optically transparent. In this method, typically a precursor molecule, such as mercaptopropyltrimethoxysilane (MPTMOS), is added to the QDs (which replaces the hydrophobic surface chains of the QD). The thiol groups of the MPTMOS react with the inorganic surface of the QD, and the methoxysilane groups polymerize through the formation of siloxane bonds, thus generating a highly crosslinked protective shell around the QD. An additional advantage of this approach is that silica exhibits a high degree of biocompatibility, and it is a simple process to functionalize its surface with appropriate (bio)analyte recognition BMs. Consequently, QDs encapsulated in a silica layer are highly suitable for further bioanalytical applications [47]. The third common approach to achieve water stabilization of QDs is to transfer the nonpolar QDs into an aqueous media combined with the use of amphiphilic polymers. Here, the hydrophobic shell of QDs (e.g. TOP [trioctylphosphine]/ TOPO) interacts with the hydrophobic alkyl chains of the amphiphilic polymeric structures through hydrophobic or electronic interactions. The hydrophilic groups of the amphiphilic polymer used will then remain oriented to the external part of the QD surface, thereby providing the required water stability [48]. A large number of amphiphilic copolymers are available for use in this approach, such as polymaleic anhydride [49]) and polyelectrolytes (poly-acrylamide [50], or biopolymers such as DNA [51]). Application of a polyethylene glycol (PEG)-based coating to QDs is another alternative often used to provide stability and biocompatibility to the NPs, Reprinted from the journal

141

13



Topics in Current Chemistry (2020) 378:35

although QDs coated with PEG spacers have reduced nonspecific protein binding, which may often limit the applicability of these NPs in bioanalytical methodologies. For such uses, PEG molecules should be previously activated with appropriate functional groups (e.g. amine, thiols or carboxyls) to provide hydrophilic bridges between the QD surface and the PEG chains [52]. A variant of this third stabilization approach consists of the encapsulation of the hydrophobic QDs in appropriate hydrophilic vehicles, such as liposomes [53]. The hollow spherical structure of liposomes and the high loading capacity makes them attractive carriers for hydrophobic QDs. Moreover, the surface of liposomes can be easily modified with the appropriate functional groups so as to allow a simple further bioconjugation with proteins, thereby minimizing nonspecific interactions of water-soluble and water-insoluble QDs with surface material and amplifying the analytical signal due to the possibility to incorporate several QDs in a single nanoliposome. In this context, a signal amplification platform based on the measurement of fluorescence from QDs encapsulated in liposomes has been recently proposed for the highly sensitive detection of human telomerase activity [54]. In the approach described, similar to a typical hybridization bioassay, biotinylated liposomes containing the QDs were recognized by a capture probe conjugated with streptavidin. In a final step, the QD-encapsulated liposomes were disrupted by the controlled addition of Triton X-100, and the fluorescence intensity of the released QDs was measured to detect telomerase activity [54]. Liposomes containing hydrophobic QDs have also been employed for tumor imaging applications through the specific recognition of aptamers conjugated to the surface of the liposomes (see Fig. 5) [55]. However, a significant drawback of liposomes is their low stability when entering in  vivo media. Additionally, QDs stabilized by this approach have a substantially increased hydrodynamic diameter, which could limit their application in bioimaging and targeting procedures. It must also be taken into account that the procedure selected for hydrophilization of the QDs likely affects their suitability in subsequent bioconjugation and future bioanalytical applications. For example, some approaches can significantly increase the hydrodynamic ratios of the NPs, which in turn can lead to non-specific binding or reduced accessibility to some targets. In this context, ligand exchange provides QDs with a small hydrodynamic diameter but also with lower photoluminescence quantum yields, while encapsulation results in larger nanoprobes sizes with higher quantum yields [56].

4 QD Bioconjugation Strategies One of the main challenge to the use of QDs in biomedical applications can be considered to be the generation of robust bonding between the appropriate target recognition molecule and the surface of the NP, as this bonding will be a key parameter affecting the direct application of the QDs in biological media. In this section, we summarize some of the most relevant physicochemical processes used to attach BMs to the QD surface, a process referred to as bioconjugation.

13

142

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



Fig. 5  a Schematic illustration of the synthesis of aptamer-conjugated liposomes containing QDs and an aptamer (Apt-QLs) against the epidermal growth factor receptor. b Liposomes containing QDs were visualized by transmission electron microscopy. c Fluorescence emission was verified (excitation wavelength [λex]  550  nm, emission wavelength [λem]  620  nm). d Dynamic light scattering revealed generation of 165-nm-diameter liposomes. Reprinted from Kim et al. [55]

All NPs have a very high surface-to-volume ratio; consequently, the role of the NP surface is of paramount importance [57]. The properties of the NP surface are determined not only by their own chemical nature, but also by the layer of capping molecules, referred to as ligands because they bind to the QD surface metals in a similar way as ligands do to the central atom in the metal complexes. Thus, ligands “stop particles aggregating, resist nonspecific adsorption of surrounding molecules, and provide a conjugation point for functional biomolecules” [58]. The synthetic method utilized to prepare QDs determines not only the size, shape and chemical nature of the QDs, but also the ligands that cap their surface. In fact, these ligands are chosen mainly because they have to control the size, shape and polydispersity of the QDs during the synthesis and to maintain their homogeneous dispersion in the solvent post synthesis. Ligand exchange reactions extend the versatility of QD material.

Reprinted from the journal

143

13



Topics in Current Chemistry (2020) 378:35

The main elements of the bioconjugation process requiring attention have been described in a comprehensive article written by Sapsford et al. [56] and comprise: (1) Control over the BM/QD ratio. The desirable ratio varies with the type of the application. It should be noted that QDs are usually larger than BMs, with the possible exception of some large proteins. It should also be noted that the reactions are interfacial in nature and that such interfaces inherently polydisperse across a population of QDs (2) Control over the orientation of the BM on the QD. Optimal activity of both the QD and the BM should be maintained. (3) Control over the separation between QD and BM. This point is crucial if the platform QD–BM is to be used in a Förster resonance energy transfer (FRET) experiment [59] (4) Control over the strength of the QD–BM bond. Most clinical or in vivo experiments require permanent and, therefore, strong linkage. Many different approaches can be used to immobilize BMs onto the QD surface, with the simplest method to link a BM to a QD surface being adsorption. However, in adsorption, the attachment of a BM is rather tenuous and maintained by weak interactions only, such as hydrogen bonding, London dispersion and Coulombic forces and lone-pair electrons [60]. As an example, the proteins present in the human body tend to bind nonspecifically onto the surface of QDs, a process which is to be avoided. A closely related approach is based on pure electrostatic interactions between the BM and the QD [61]. However, while this method of functionalization is generally straightforward and fast, electrostatic interaction, similar to simple adsorption, suffers from serious disadvantages, including instability, lack of orientation control on the BM and on the BM/QD ratio [62, 63]. Therefore, the most common routes to achieve bioconjugation consist of the four shown in Fig. 6 [56]. 4.1 Direct Union of BMs BMs can be joined to a semiconductor QD surface through the direct covalent union of the BMs to the surface of the QD semiconductor (route 1 in Fig. 6). Thus, proteins, peptides and nucleic acids can be bonded to the QD surface metal atoms (especially Zn) through their (cysteine) thiol and (histidine) imidazole groups [56]. Thiol and histidine motifs could eventually be added to the “natural” BMs and, occasionally, thiol groups are created, reducing peripheral S–S bonds. Occasionally, BMs are bonded directly to the QD in the synthesis processes (biological templating) [56]. 4.2 BMs Bonded to a Ligand A more general method to join BMs to a QD surface is to bind covalently the BM to a ligand that has been previously attached to the QD surface (route 2 in Fig. 6). It should be pointed out that GQDs usually contain carboxyl, hydroxyl, carbonyl and epoxide external groups [64] and that carbon-based QDs (CQDs) can be synthesized

13

144

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35 Native ligands L

BM BM

L 4

bioconjugated QD

native QD

3 L

1

L

The four main routes to bioconjugugation

BM 1: 2: 3: 4:

2 L

BM

direct union BM bonded to a ligand L BM bonded to an encapsulating shell BM bonded through Biotin/Sreptavidin

Fig. 6  The four main routes to join a biomolecule (BM) to a semiconductor QD surface. (1) Through a direct covalent bond, (2) through a covalent bond between the BM and a ligand (L) that was previously anchored onto the QD surface, (3) covalent binding of the BM to a terminal functional group integrated in an encapsulating shell; (4) taking advantage of the specific union biotin/streptavidin

to have selectively carboxylic, amine or other groups in their surface [65]. On the contrary, native semiconductor QDs, with some exceptions being those prepared in water [66, 67], do not have the appropriate ligands able to be robustly bonded to a BM. However, this is not a problem because, as seen in the preceding section, the original ligands can be replaced by others. The new ligands could be monodentate or bidentate simple molecules, such as 3-mercaptopropionic acid (MPA) or dihydrolipoic acid (DHLA), respectively [68], that are attachable to the QD surface through the thiol group(s) and which possess a free terminal carboxylic group to join the BM. However, the new ligands are usually complex molecules with an anchoring group(s) (e.g. polythiol), a spacer chain (frequently a hydrophilic segment) and a terminal functional group (carboxylic or amine groups among others) [69]. Very often, the spacer chain is PEG, that is the whole ligand is a bifunctional PEG molecule [67] because PEG is biocompatible and highly soluble in water and stabilizes QDs against aggregation. 4.3 BMs Bonded to an Encapsulating Shell The native ligand exchange method is associated with a number of problems, such as a relatively weaker bond between the thiol group and the metal of the QD surface, reduced photoluminescent quantum yield, among others [61, 63]. An alternative is the encapsulation of the QDs (route 3 in Fig. 6), either with a layer of amorphous silica or with a copolymer. The formation of a silica outer sphere (silanization) increases the solubility and stability of the QDs and retains most of the emission properties [63]. A variety of silanization processes have been described [63, 67, 70, 71]. In general, all are laborious and require several steps [63], including bonding of the silica layer onto the QD surface through an anchoring group. The most commonly used anchoring groups are Reprinted from the journal

145

13



Topics in Current Chemistry (2020) 378:35

thiol or amino groups (A in Fig. 7a). The silica layer which also possess functional groups at the periphery, such as thiol and amino, among others (F in Fig. 7a). The encapsulation of a semiconductor QD with an amphiphilic copolymer is possible due to the ability of the latter’s long hydrophobic tails to interact and interdigitate with the pristine QD ligands (such as TOP, TOPO, hexadecylamine, stearic acid, etc.) [67], thereby leaving the hydrophilic and functionalized segments in contact with the water molecules of the solvent [61, 63, 70]. Although the copolymers of maleic anhydride or acrylic acid are the most popular [61, 63, 70, 72, 73], there are many others (e.g. phospholipid–PEG copolymers, which possess a terminal functional group [74]) (see Fig. 7b). Some authors consider the envelopment of several QDs inside a polyethylenimine coat [75] or inside the bilayer of a liposome [76, 77], as specific cases of QD encapsulation. Although these examples are not strictly comparable to those described in Fig.  7, in which every QD is singularly encapsulated, there is certainly some similarity. 4.4 BMs Bonded Through Biotin/Streptavidin Another method used very frequently to join BMs to the QD surface is based on the strong interaction between biotin and avidin, with a dissociation constant of ­10−15 M [78]. Avidin, a protein found in egg white, contains four identical subunits, each with a single biotin-binding site. Biotin is vitamin H. In this method (route 4 in Fig. 6), it is considered advantageous to substitute avidin with deglycosylated avidin derivatives, such as streptavidin or neutravidin [56, 62]. There are a variety of biotin

(a)

(b)

SILICA SHELL

COPOLYMER ENCAPSULATION COOH

COOH

Phospholipides F-PEG

QUANTUM DOT

Si O Si O O A Si O Si O O Si O Si

F

COOH

QUANTUM DOT

F

COOH

COOH PAA and PMA copolymers

F F

native ligands

Fig. 7  Encapsulation of a semiconductor QD with silica shell (a) or with copolymers (b). The silica layer binds to the QD surface through an anchoring group (A) and possesses functional groups (F) at the periphery. The copolymers most frequently used for encapsulation are polymaleic anhydride (PMA)- or poly(acrylic acid) (PAA)-based copolymers, but end-functionalized polyethylene glycol (F-PEG) phospholipids are also used

13

146

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35

derivatives that make the biotinylation of QDs and BMs a rather straightforward process [56, 62]. QDs and BMs may be also functionalized with avidin, although it should be noted that the attachment of avidin to the QD during initial modification will probably obscure one or more available biotin binding sites [56]; at a later moment, however, avidin-functionalized QDs could be attached to biotinylated BMs. Conversely, biotin-functionalized QDs can be used to join avidin-functionalized BMs. Even biotin-labeled QDs can be coupled to biotin-labeled BMs through an intermediary avidin linker [56, 61] although unexpected results may be obtained since the avidin can bind up to four biotin moieties [56, 61]. There are many more types of specific non-covalent affinity between pairs of molecules other than avidin/biotin interactions that are useful in bioconjugation; these include histidine–nickel nitrilotriacetic acid interaction, barnase–barstar interaction and antibody–ligand interaction [62]. 4.5 Covalent Coupling Strategies

O O

P

O

O

O C

1

8

H 7

CO OH

The focus in this section is on covalent bonding between ligands anchored onto the QD surface (directly or attached to an encapsulating shell) and the incoming BMs. In this regard, we consider biotin and avidin (or straptividin or neutravidin) to be BMs. The number of functional groups useful for bioconjugation reactions is rather limited in natural BMs [56] (see Fig. 8). Carboxylic and amino groups are present in peptides and proteins not only as terminal groups but also as side groups in peptides such as aspartic or glutamic acids or lysine, respectively. Less common in peptides and proteins are the thiol, phenol and imidazole groups that can be found in the amino acids cysteine, tyrosine and histidine, respectively. Other functional groups

Naturally ocurring biomolecules

6 HO

5

NH

2

3

SH

CARBOHYDRATES 6.- Alcohol 7.- Aldehyde

4

N

OH

N H

2

AMINOACIDS 1.- Carboxylic (Glutamic, Aspartic) 2.- Amino (Lysine) 3.- Thiol (Cysteine) 4.- Phenol (Tyrosine) 5.- Imidazole (Hystidine)

NUCLEOTIDES 8.- Phosphate

Fig. 8  The most important functional groups found in naturally occurring BMs. 1–5 Functional groups present in amino acids: carboxylic, amino, thiol, phenol, imidazole. 6, 7 Functional groups from carbohydrates and derivatives: alcohol and aldehyde. 8 Functional group derived from nucleotides: phosphate Reprinted from the journal

147

13



Topics in Current Chemistry (2020) 378:35

that are available in even less common amino acids, such as tryptophan, will not be considered in this review. Carbohydrates and their derivatives provide alcohol and aldehyde as reactive groups, the latter are obtained by oxidation of the former. Nucleic acids possess sugars, phosphates and some bases (the bases are not included in Fig. 8 because they are not usually modified). Although the number of available functional groups in BMs for bioconjugation reactions would appear to be low, the number of possible reactions by which BMs can be joined covalently to the ligands attached to the QDs described in the literature and used in the commercial sources is rather numerous [56, 62, 63, 71, 76, 79–81]. However, only a small portion of these seem to have been actually used when QDs were involved [62, 63, 82]. One of the most well-studied and easy-to-perform bioconjugation reactions is between a terminal carboxylic acid and a peripheral amine group, conducted under mild conditions, to yield an amide group, with the help of EDC (1-ethyl-3-(3dimethylaminopropyl) carbodiimide hydrochloride) and sulfo-NHS (N-Hydroxysulfosuccinimide sodium salt). Despite all the disadvantages that this method presents, it is still widely used [62, 63, 82]. An alternative is to use carbonyldiimidazol (CDI) instead of the EDC/sulfo-NHS pair [63, 83]. Other common routes are the reaction of amines with carbonyl groups to yield imine groups, which are usually subsequently reduced with sodium cyanoborohydride [62, 63], and the Michael addition of a terminal thiol to a maleimido group [63, 80]. Another popular coupling method is the utilization of heterobifunctional molecules, such as sulfosuccinimidyl-4-(N-maleimidomethyl) cyclohexane-1-carboxylate (sulfo–SMCC). The NHS ester end of sulfo-SMCC can react with primary amine groups, and the other terminal maleimido function can add a thiol [62, 63]. Similar crosslinker molecules are described in the literature [56, 62, 80]. 4.6 Bioorhogonality The traditional coupling methods described in this review thus far have a number of limitations, among which the most important is undesirable side reactions [56]. The solution to this and other problems is bioorthogonal chemistry. Bioorthogonal chemical reactions involve only the target functions (in QDs and BMs) and do not affect the other functional groups present in either the affected QD and BM or in the biological environment [84, 85]. Although numerous biorthogonal chemical reactions are described in the literature [84–86], only some of these seem to have been carried out when QDs are involved (see those described in Fig.  9). The latter include the copper-catalyzed alkyne-azide cycloaddition (click chemistry); the cycloaddition between tetrazine and strained double bonds (tetrazine ligation); and hydrazone formation by reacting hydrazine and carbonyl groups (hydrazine ligation). It should be noted that not one of the functional groups shown in Fig.  8, i.e those present in “natural” BMs, is involved in these bioorthogonal reactions. This is, of course, the most important advantage of bioorthogonal chemical reactions: the reactions, as described in Fig. 9, do not affect normal molecules present in the biological milieu. Some effort would

13

148

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35 "CLICK" CHEMISTRY Cu+

N

N3

N N

Alkyne + Azide (Cu+)

TETRAZINE LIGATION N N N N Norbornene

+

Tetrazine

HYDRAZONE LIGATION O

O

N

N H

O

NH2

Hydrazine

N +

N H

N

Carbonyl

Fig. 9  The main bio-orthogonal reactions carried out utilizing QDs. Top: copper-catalyzed alkyne-azide cycloaddition. Middle: cycloaddition between tetrazine and strained double bonds. Bottom: hydrazone formation by reacting hydrazine and carbonyl groups

need to be made to design procedures for attaching these “new” functional groups to the natural BMs, however no difficulties are foreseeable to bind these to the QDs. In this regard, it has been determined that virtually any functional group can be sitespecifically introduced into peptides and nucleotides as needed during the initial synthesis or through subsequent modification [56].

5 QD Biosensing Applications: Point‑of‑Care Diagnostics Quantum dot–BM hybrids (i.e. those bound to antibodies, DNA, aptamers, etc.) have been used as sensing probes in a wide variety of in vitro bioassays and biosensors for the detection of different clinical relevant BMs. QDs can be used as labels in a wide spectrum of detection methods (Fig. 10). Most of the reported QD-based bioassays and biosensors have been developed by using QDs as fluorescence labels (Fig.  10a) in fluorescence quenching-based (turn-off), fluorescence enhancementbased (turn-on) and—especially—FRET assays [87–92]. FRET is a very sensitive technology for studying BM interactions that involves the transfer of energy from Reprinted from the journal

149

13



Topics in Current Chemistry (2020) 378:35

Fig. 10  Detection methods for QD-based biosensing. a Fluorescence. Jablonski diagram explaining the effect of fluorescence (left), and a photograph and emission spectra illustrating size-controlled fluorescence of QDs. Adapted from Wen et  al. [87], with permission. b Förster resonance energy transfer (FRET): Jablonski diagram explaining the effect of FRET (left) and schematic illustration of a FRETbased sandwich bioassay (right). c Electrochemiluminescence (ECL): schematic illustration explaining anodic ECL. A hole is created by the electrode in the valence band of QDs with the concomitant injection of an electron from a previously oxidized coreactant (C). The recombination of electron and hole lead to an anodic ECL emission. d Photoelectrochemical (PEC) reaction: schematic illustration explaining anodic PEC reaction. An electron–hole pair is created on QDs after their photoexcitation. The electron transferred from the valence band to the conduction band of the photoexcited QDs is then ejected to the electrode with the concomitant transfer of electrons from an electron donor (C), generating an anodic photocurrent. e Electrochemical (EC) reaction: schematic illustration explaining the electrochemical detection of QDs by anodic stripping voltammetry. After the dissolution of QDs, metallic species are deposited (reduced) on the electrode and re-oxidized again to be detected Adapted from from Wen et al. [87], copyright 2017, with permission

an excited state donor (usually a fluorophore) to a proximal ( 100-fold when NIR QDs emitting at 1320  nm (NIR-II region) were employed as contrast agents rather than QDs emitting at 850 nm (NIR-I region), with reduced autofluorescence and a superior tissue penetration observed at the longer wavelengths. Thus, more recently there have been some efforts to try to develop QDs that emit in the NIR-II window, based on NP cores of different metallic elements, such as PbSe, PbS or hybrid CdHgTe. More recently, NIR-II QDs have also been synthesized that avoid the use of toxic elements: a mirage of reports propose ­Ag2S QDs as NIR-II emitting agents [154].

13

164

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



To date, there have been only a few reports on successful specific targeting employing ­Ag2S QDs [23]. In all such reportes, ­Ag2S NPs were surface functionalized with appropriate recognition elements for in vivo imaging following different bioconjugation approaches. In a very recent study, A ­ g2S NPs were surface functionalized with plerixafor, a small molecule drug used for the inhibition of CXC chemokine receptor 4 (CXCR4). The bioconjugate was used for in  vivo imaging of metastatic breast cancer cells based on the selective linkage of the functionalized ­Ag2S NPs to highly metastatic breast cancer cells (4T1 tumor model) via their CXCR4 receptor [155]. Moreover, the use of the surface-decorated ­Ag2S NPs together with their photothermal properties resulted in a unique and tumor-specific theranostic element. 6.3 Multimodal Imaging There is currently an increasing interest among researchers on the development of new nanomaterials for multimodal imaging applications in biology and medicine. In this context, multimodal fluorescent-magnetic based nanomaterials deserve particular attention as they can be used both as diagnostic and drug delivery tools, which could facilitate the diagnosis and treatment of many diseases. The frequently investigated QD-based hybrid-NPs with multiple capabilities are probably the magnetic-QDs. As an example, differently sized infrared-emitting QDs have been incorporated, together with variable amounts of Fe-based magnetic NPs, into poly(styrene/acrylamide) copolymer nanospheres for the preparation of fluorescent-magnetic nanocomposites [156]. The dual-encoded nanobioprobes developed as such exhibited different luminescent behavior (due to the different sizes of the NPs) and magnetic susceptibility. They were proven to be capable of simultaneously recognizing and separating multiple biocomponents from complex samples when three kinds of lectins were used as the targets. Multimodal imaging can integrate structural/functional information from several imaging modalities, thus promising more accurate diagnosis than any single imaging modality. One important advantageous feature of liposome encapsulation is the possibility to co-immobilize several NPs exhibiting different properties to develop multimodal imaging platforms. In a very recent article, Xu et al. reported the integration of a theranostic liposome (QSC-Lip) with superparamagnetic iron oxide NPs (SPIONs) and QDs and cilengitide (CGT) into one platform, with the aim to target glioma in magnetic targeting (MT) for guiding the surgical resection of glioma [157]. In vivo dual-imaging studies show that QSC-Lip not only produces an obvious negative-contrast enhancement effect on glioma by MRI but also makes tumor emitting fluorescence under MT. Among the different techniques widely used for molecular imaging, MRI is currently one of the main in  vivo imaging techniques used routinely in diagnosis, while fluorescence imaging is nowadays most widely used for in vitro studies; thus these two imaging techniques are complementary. Clearly, there is much research interest directed towards preparing fluorescent imaging/MRI imaging dual‐modality nanoprobes to be used in many diagnostic and biomedical applications, such Reprinted from the journal

165

13

Topics in Current Chemistry (2020) 378:35

as cell labeling, enzyme activity measurements, tumor diagnosis and therapy and anatomical localization and real‐time assessment during surgery [158]. Nanostructures based on fluorescent QDs can be synthesized to provide magnetic properties to the nanomaterial, thereby creating opportunities for multi-modality biomedical imaging. Fluorescent QDs exhibiting magnetic susceptibility can be synthesized following four different methodologies: metal doping, covalent conjugation, isocrystal growth and co-encapsulation or electrostatic assembly. In particular, a large number of fluorescent imaging/MRI dual-modality imaging nanoprobes combine G ­ d3+ or 2+ ­Mn ions with QDs. As an example, a dual contrast nanoreagent was developed by doping Gd ions into ­CuInS2/ZnS QDs (Fig. 18) [159]. The resulting NPs exhibited NIR fluorescence emission and MRI contrast capabilities with a high longitudinal relaxivity (r1), which was 2.5-fold higher than that of clinically approved Gd agents. In addition, the in vivo imaging experiments showed that the Gd-doped NPs could enhance both NIR fluorescence and T1-weighted MRI of tumor tissue through passive targeting accumulation.

7 Conclusions and Perspectives In general, photoluminescent QD probes are widespread and used in countless biomedical applications. For biosensing, a great potential of QD-conjugates also lies in multiplexing as well as in vitro and in vivo fluorescent imaging. There has been major progress in the development of in  vivo drug delivery systems, and interest remains high in this area. It is important to note that all of these applications are possible because of the advances in QD stabilization in biological media by appropriate surface functionalization and their bioconjugation to suitable BMs.

Fig. 18  Fabrication procedure and functional description of the Gd-doped QDs with dual-mode imaging capabilities. Reprinted from Yang et al. [159], copyright 2017, with permission from the American Chemical Society

13

166

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



All of the recent advances in bioconjugation chemistry has made it possible to attach almost any BM of interest to the surface of a QD. However, work is till needed to further enhance the development and applications of QD bioconjugates. In this context, we have identified the following issues. The orientation of the BMs is a key issue that still needs to be addressed. Many often random orientations are sufficient when the QDs are used in conventional hybridization applications. However, controlled orientations may be needed for the assembly of functionalized 3D structures. The functionalization of QDs with different BMs (e.g. antibodies, peptides, nucleic acids or aptamers) offer a wide range of opportunities for applying the nanoassemblies in clinical diagnosis, including ultrasensitive detection of disease biomarkers, in  vivo targeted imaging or drug delivery applications. The multiplexing capabilities of these QDs open new avenues for the use of differently sized QDs for the simultaneous detection of multiple biomarkers (a key aspect in efficient clinical diagnosis). Moreover, encapsulation of multiple QDs in highvolume nanocarriers (e.g. nanosomes or PLGA NPs) may enable the construction of a panel of multifunctional systems for targeted drug delivery and molecular imaging. The development of new multimodal imaging nanoprobes is a focus of many researchers. The use of NPs as imaging probes offers several advantages over conventional molecular-scale contrast agents; these include high loading capacity, where the concentration of the imaging agents can be controlled within each NP during the synthesis process; tunable surface that can potentially extend the circulation time of the contrast agents in the blood or target them to specific locations in the body; or provision of multimodal imaging capacities because NPs can combine two or more contrast properties, which can be used in multiple imaging techniques simultaneously [160]. Recent advances in nanotechnology has enabled the development of multifunctional QDs by doping the nanocrystals with appropriate metals, thus integrating two or more imaging contrast agents and thereby enabling their detection by different imaging techniques [42]. One of the major challenges when developing novel bioassay methods for clinical applications is the requirement for high sensitivity in the detection because of the ultralow concentrations of the biomarkers to be detected. Here, the use of QDs as tags in immunoassays could be a powerful approach to achieve the desired ultrasensitivity. As an example, ultrahigh sensitivity for BMs could be easily achieved through metal deposition on the surface of the NP tags acting as catalytic seeds, thus effectively amplifying the size of the metallic NPs after the immunoassay [161]. Obviously, QDs cannot be safely used as tags for in  vivo applications until the problem of their toxicity is solved. Despite very extensive studies of toxicity of QDs in different cellular and animal models, the in vivo toxicological effect of QDs remains controversial [37]. The possible release of toxic heavy metals from the core of the QDs as a result of intensive UV illumination has to be taken into account [44]. The preparation of heavy metal-free QDs is being addressed as a promising avenue to overcome such toxicity problems. Additionally, it must be considered that an ideal solubilization strategy should reduce QD toxicity and undesirable nonspecific QD uptake by living tissues, thus reducing cytotoxic effects. Reprinted from the journal

167

13



Topics in Current Chemistry (2020) 378:35

Paper-based microfluidic systems have been revealed as the most suitable platform for POC analysis, with the use of QDs as labels becoming increasingly popular in the development of this type of systems. Although fluorescence QD-based POC systems using hand-held readers or even smartphone-based detectors have been successfully reported, they suffer from an important limitation related to the need for bulky and complex detectors for quantification. However, it is expected that the use of alternative detection methods (e.g. electrochemical) and the rapid development of portable devices and mobile phone technology will allow the miniaturization of the detection systems for POC devices in the near future. Miniaturized signal-recording devices also require a merging of QD barcode technology and POC testing. In the sensing field, there is also a great expectation for recently developed GQDs and CQDs. To summarize, even though there is still a long road to go before bioconjugated QDs are considered to be routine in in vitro and especially in vivo diagnosis, overall we firmly believe that the rapid development of new bioconjugated nanomaterials will move bioconjugated QDs forward to real-life diagnostic applications in modern biology and medicine. Acknowledgements Financial support from the FC-GRUPIN-ID/2018/000166 project (Asturias Regional Government, Spain) and the CTQ2017–86994-R and CTQ2016–79412-P projects (MINECO, Spain) is gratefully acknowledged. A. de la Escosura-Muñiz acknowledges the MICINN (Spain) for the “Ramón y Cajal” Research Fellow (RyC-2016-20299).

References 1. Prado M, Espiña B, Fernández-Argüelles MT, Diéguez L, Fuciños P, Vial S, Oliveira JM, Reis RL, Boehme K (2016) Detection of foodborne pathogens using nanoparticles. Advantages and trends. In: Barros-Velázquez J (ed) Antimicrobial food packaging. Elsevier, Amsterdam, pp 183–201 2. Trapiella-Alfonso L, Llano-Suárez P, Sanz-Medel A, Costa-Fernández JM, Fernández-Argüelles MT (2017) Analytical nanoscience and nanotechnology. In: Meyers RA (ed) Encyclopedia of analytical chemistry. Wiley, Hoboken, pp 1–24 3. Lopez-Lorente A, Valcarcel M (2016) The third way in analytical nanoscience and nanotechnology: involvement of nanotools and nanoanalytes in the same analytical process. Trends Anal Chem 75:1–9 4. Blanco-López MC, Rivas M (2019) Nanoparticles for bioanalysis. Anal Bioanal Chem 411:1789–1790 5. Ahsan MA, Jabbari V, Imam MA, Castro E, Kim H, Curry ML, Valles-Rosales DJ, Noveron JC (2020) Nanoscale nickel metal organic framework decorated over graphene oxide and carbon nanotubes for water remediation. Sci Total Environ 698:134214 6. Cova CM, Zuliani A, Santiago ARP, Caballero A, Muñoz-Batista MJ, Luque R (2018) Microwaveassisted preparation of Ag/Ag2S carbon hybrid structures from pig bristles as efficient HER catalysts. J Mater Chem A 6:21516–21523 7. Ahsan MA, Deemer E, Fernandez-Delgado O, Wang H, Curry ML, El-Gendy AA, Noveron JC (2019) Fe nanoparticles encapsulated in MOF-derived carbon for the reduction of 4-nitrophenol and methyl orange in water. Catal Commun 130:105753 8. Ahsan MA, Jabbari V, El-Gendy AA, Curry ML, Noveron JC (2019) Ultrafast catalytic reduction of environmental pollutants in water via MOF-derived magnetic Ni and Cu nanoparticles encapsulated in porous carbon. App Surf Sci 497:143608 9. Ahsan MA, Fernandez-Delgado O, Deemer E, Wang H, El-Gendy AA, Curry ML, Noveron JC (2019) Carbonization of Co-BDC MOF results in magnetic C@Co nanoparticles that catalyze the reduction of methyl orange and 4-nitrophenol in water. J Mol Liq 290:111059

13

168

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



10. Sanad MF, Shalan AE, Bazid SM, Serea ESA, Hashem EM, Nabiha S, Ahsan MA (2019) A graphene gold nanocomposite-based 5-FU drug and the enhancement of the MCF-7 cell line treatment. RSC Advances 9:31021–31029 11. Dominguez N, Torres B, Barrera LA, Rincon JE, Lin Y, Chianelli RR, Ahsan MA, Noveron JC (2018) Bimetallic CoMoS composite anchored to biocarbon fibers as a high-capacity anode for Liion batteries. ACS Omega 3:10243–10249 12. Piccinno F, Gottschalk F, Seeger S, Nowack B (2012) Industrial production quantities and uses of ten engineered nanomaterials in Europe and the world. J Nanopart Res 14:1109 13. Snee PT (2020) Semiconductor quantum dot FRET: untangling energy transfer mechanisms in bioanalytical assays. Trends Anal Chem 123:115750 14. Llano Suarez P, García-Cortés M, Fernández-Argüelles MT, Ruiz Encinar J, Valledor M, Ferrero FJ, Campo JC, Costa-Fernandez JM (2019) Functionalized phosphorescent nanoparticles in (bio) chemical sensing and imaging—a review. Anal Chim Acta 1046:16–31 15. Granada-Ramírez DA, Arias-Cerón JS, Rodriguez-Fragoso P, Vázquez-Hernández F, Luna-Arias JP, Herrera-Perez JL, Mendoza-Álvarez JG (2017) Quantum dots for biomedical applications. In: Nanobiomaterials: nanostructured materials for biomedical applications. Unidad Profesional Interdisciplinaria de Ingeniería y Tecnologías Avanzadas (UPIITA), pp 411–436. https​://doi. org/10.1016/B978-0-08-10071​6-7.00016​-7 16. Coto-García AM, Sotelo-González E, Fernández-Argüelles MT, Pereiro R, Costa-Fernández JM, Sanz-Medel A (2011) Nanoparticles as fluorescent labels for optical imaging and sensing in genomics and proteomics. Anal Bioanal Chem 399:29–42 17. Geißler D, Charbonnière LJ, Ziessel RF, Butlin NG, Löhmannsröben H-G, Hildebrandt N (2010) Quantum dot biosensors for ultrasensitive multiplexed diagnostics. Angew Chem 49:1396–1401 18. Hildebrandt N (2011) Biofunctional quantum dots: controlled conjugation for multiplexed biosensors. ACS Nano 5:5286–5290 19. Chan WC, Nie S (1998) Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science 281:2016–2018 20. Sukhanova A, Devy J, Venteo L, Kaplan H, Artemyev M, Oleinikov V, Klinov D, Pluot M, Cohen JH, Nabiev I (2004) Biocompatible fluorescent nanocrystals for immunolabeling of membrane proteins and cells. Anal Biochem 324:60–67 21. Gill R, Zayats M, Willner I (2008) Semiconductor quantum dots for bioanalysis. Angew Chem Int Ed 47:7602–7625 22. Bear J, Charron G, Fernandez-Arguelles MT, Massadeh S, McNaughter P, Nann T (2011) In vivo applications of inorganic nanoparticles. In: Booss-Bavnbek B, Klosgen B, Larsen J, Pociot F, Renstrom E (eds) BetaSys. Systems biology, vol 2. Springer, New York, pp 185–220 23. Lu C, Chen G, Yu B, Cong H (2018) Recent advances of low biological toxicity A ­ g2S QDs for biomedical application. Adv Eng Mater 20:1700940 24. Rabani E (2001) Structure and electrostatic properties of passivated CdSe nanocrystals. J Chem Phys 115:1493 25. Jin Z, Hildebrandt N (2012) Semiconductor quantum dots for in  vitro diagnostics and cellular imaging. Trends Biotechnol 30:394–403 26. Alivisatos AP (1996) Perspectives on the physical chemistry of semiconductor nanocrystals. J Phys Chem 100:13226–13239 27. Coto-Garcia AM, Fernandez-Arguelles MT, Costa-Fernandez JM, Sanz-Medel A, Valledor M, Campo JC, Ferrero FJ (2013) The influence of surface coating on the properties of water-soluble CdSe and CdSe/ZnS quantum dots. J Nanopart Res 15:1330 28. Fernandez-Arguelles MT, Costa-Fernandez JM, Pereiro R, Sanz-Medel A (2010) Organically modified quantum dots in chemical and biochemical analysis. In: Martinez-Manez R, Rurack K (eds) The supramolecular chemistry of organic–inorganic hybrid materials. Wiley, Hoboken, pp 377–403 29. Wagner MK, Li F, Li J, Li X, Le XC (2010) Use of quantum dots in the development of assays for cancer biomarkers. Anal Bioanal Chem 397:3213–3224 30. Garcia-Cortes M, Sotelo E, Fernandez-Arguelles MT, Encinar JR, Costa-Fernandez JM, SanzMedel A (2017) Capping of Mn-doped ZnS quantum dots with DHLA for their stabilization in aqueous media: determination of the nanoparticle number concentration and surface ligand density. Langmuir 33:6333–6341 31. Wu P, Yan XP (2013) Doped quantum dots for chemo/biosensing and bioimaging. Chem Soc Rev 42:5489–5521 Reprinted from the journal

169

13



Topics in Current Chemistry (2020) 378:35

32. Panda SK, Hickey SG, Demir HV, Eychmüller A (2011) Bright white-light emitting manganese and copper co-doped ZnSe quantum dots. Angew Chem Int Ed 50:4432–4436 33. Ehlert O, Osvet A, Batentschuk M, Winnacker A, Nann T (2006) Synthesis and spectroscopic investigations of Cu- and Pb-doped colloidal ZnS nanocrystals. J Phys Chem B 110:23175–23178 34. Sotelo-Gonzalez E, Fernandez-Arguelles MT, Costa-Fernandez JM, Sanz-Medel A (2012) Mndoped ZnS quantum dots for the determination of acetone by phosphorescence attenuation. Anal Chim Acta 712:120–126 35. Mohapatra S, Rout SR, Das RK, Nayak S, Ghosh SK (2016) Highly hydrophilic luminescent magnetic mesoporous carbon nanospheres for controlled release of anticancer drug and multimodal imaging. Langmuir 32:1611–1620 36. Wang J, Zhang Z, Zha S, Zhu Y, Wu P, Ehrenberg B, Chen JY (2014) Carbon nanodots featuring efficient FRET for two-photon photodynamic cancer therapy with a low fs laser power density. Biomaterials 35:9372–9381 37. Chong Y, Ma Y, Shen H, Tu X, Zhou X, Xu J, Dai J, Fan S, Zhang Z (2014) The in  vitro and in vivo toxicity of graphene quantum dots. Biomaterials 35:5041–5048 38. Wang J, Zhang P, Huang C, Liu G, Leung KCF, Wáng YXJ (2015) High performance photoluminescent carbon dots for in vitro and in vivo bioimaging: effect of nitrogen doping ratios. Langmuir 31:8063–8073 39. Vasimalai N, Vilas-Boas V, Gallo J, Cerqueira MF, Menéndez-Miranda M, Costa-Fernández JM, Diéguez L, Espiña B, Fernández-Argüelles MT (2018) Green synthesis of fluorescent carbon dots from spices for in vitro imaging and tumour cell growth inhibition. Beilstein J Nanotechnol 9:530–544 40. Righetto M, Privitera A, Fortunati I, Mosconi D, Zerbetto M, Curri ML, Corricelli M, Moretto A, Agnoli S, Franco L, Bozio R, Ferrante C (2017) Spectroscopic insights into carbon dot systems. J Phys Chem Lett 8:2236–2242 41. Dong Y, Wang R, Li G, Chen C, Chi Y, Chen G (2012) Polyamine-functionalized carbon quantum dots as fluorescent probes for selective and sensitive detection of copper ions. Anal Chem 84:6220–6224 42. Bouzas-Ramos D, Cigales-Canga J, Mayo JC, Sainz RM, Ruiz Encinar J, Costa-Fernandez JM (2019) Carbon quantum dots codoped with nitrogen and lanthanides for multimodal imaging. Adv Funct Mater 29:1903884 43. Manna L, Scher EC, Li LS, Alivisatos AP (2002) Epitaxial growth and photochemical annealing of graded CdS/ZnS shells on colloidal CdSe nanorods. J Am Chem Soc 124:7136–7145 44. Coto-Garcia AM, Valledor M, Campo JC, Ferrero FJ, Fernandez-Argüelles MT, Costa-Fernandez JM, Sanz-Medel A (2011) Dynamic analysis of the photoenhancement process of colloidal quantum dots with different surface modifications. Nanotechnology 22:9 45. Wu ZY, Zhao YL, Qiu FP, Li YP, Wang SW, Yang BH, Chen L, Sun JH, Wang JY (2009) Forming water-soluble CdSe/ZnS QDs using amphiphilic polymers, stearyl methacrylate/methylacrylate copolymers with different hydrophobic moiety ratios and their optical properties and stability. Colloid Surface A 350:121–129 46. Karakoti AS, Shukla R, Shanker R, Singh S (2015) Surface functionalization of quantum dots for biological applications. Adv Colloid Interface Sci 215:28–45 47. Aubert T, Soenen SJ, Wassmuth D, Cirillo M, Van Deun R, Braeckmans K, Hens Z (2014) Bright and stable CdSe/CdS@SiO(2) nanoparticles suitable for long-term cell labelling. ACS Appl Mater Interfaces 6:11714–11723 48. Palui G, Aldeek F, Wang WT, Mattoussi H (2015) Strategies for interfacing inorganic nanocrystals with biological systems based on polymer-coating. Chem Soc Rev 44:193–227 49. Fernandez-Arguelles MT, Costa-Fernandez JM, Pereiro R, Sanz-Medel A (2008) Simple bio-conjugation of polymer-coated quantum dots with antibodies for fluorescence-based immunoassays. Analyst 133:444–447 50. Potapova I, Mruk R, Prehl S, Zentel R, Basche T, Mews A (2003) Semiconductor nanocrystals with multifunctional polymer ligands. J Am Chem Soc 125:320–321 51. Mattoussi H, Mauro JM, Goldman ER, Green TM, Anderson GP, Sundar VC, Bawendi MG (2001) Bioconjugation of highly luminescent colloidal CdSe-ZnS quantum dots with an engineered twodomain recombinant protein. Phys Status Solidi B 224:277–283 52. Du Y, Zhong Y, Donga J, Qiana C, Sunc S, Gaod L, Yang D (2019) The effect of PEG functionalization on the in vivo behavior and toxicity of CdTe quantum dots. RSC Advances 9:12218–12225

13

170

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35



53. Beloglazova NV, Shmelin PS, Speranskaya ES, Lucas B, Helmbrecht C, Knopp D, Niessner R, De Saeger S, Goryacheva IY (2013) Quantum dot loaded liposomes as fluorescent labels for immunoassay. Anal Chem 85:7197–7204 54. Zavari-Nematabad A, Alizadeh-Ghodsi M, Hamishehkar H, Alipour E, Pilehvar-Soltanahmadi Y, Zarghami N (2017) Development of quantum-dot-encapsulated liposome-based optical nanobiosensor for detection of telomerase activity without target amplification. Anal Bioanal Chem 409:1301–1310 55. Kim MW, Jeong HY, Kang SJ, Choi MJ, You YM, Im CS, Lee TS, Song IH, Lee CG, Rhee KJ, Lee YK, Park YS (2017) Cancer-targeted nucleic acid delivery and quantum dot imaging using EGF receptor aptamer-conjugated lipid nanoparticles. Sci Rep 7: 9474. https​://doi.org/10.1038/ s4159​8-017-09555​-w 56. Sapsford KE, Algar WR, Berti L, Gemmill KB, Casey BJ, Oh E, Stewart MH, Medintz IL (2013) Functionalizing nanoparticles with biological molecules: developing chemistries that facilitate nanotechnology. Chem Rev 113:1904–2074 57. Boles MA, Ling D, Hyeon T, Talapin DV (2016) The surface science of nanocrystals. Nat Mater 15:141–153 58. Howes PD, Chandrawati R, Stevens MM (2014) Colloidal nanoparticles as advanced biological sensors. Science 346:1247390 59. Hildebrandt N, Spillmann CM, Algar WR, Pons T, Stewart MH, Oh E, Susumu K, Díaz SA, Delehanty JB, Medintz IL (2017) Energy transfer with semiconductor quantum dot bioconjugates: a versatile platform for biosensing, energy harvesting, and other developing applications. Chem Rev 117:536–711 60. Xia X-R, Monteiro-Riviere NA, Riviere JE (2010) An index for characterization of nanomaterials in biological systems. Nat Nanotechnol 5:671–675 61. Zhang F, Lees E, Amin F, Rivera-Gil P, Yang F (2011) Polymer-coated nanoparticles: a universal tool for biolabelling experiments. Small 7:3113–3127 62. Foubert A, Beloglazova NV, Rajkovic A, Sas B, Madder A, Goryacheva IY, De Saeger S (2016) Bioconjugation of quantum dots: review and impact on future application”. Trends Anal Chem 83:31–48 63. Karakoti AS, Shukla R, Shanker R, Singh S (2015) Surface functionalization of quantum dots for biological applications. Adv Colloid Interface 215:28–45 64. Pedrero M, Campuzano S, Pingarrón JM (2017) Electrochemical (bio)sensing of clinical markers using quantum dots. Electroanalysis 29:24–37 65. Liu W, Li Ch, Ren Y, Sun X, Pan W, Li Y, Wang J, Wang W (2016) Carbon dots: surface engineering and applications. J Mater Chem B 4:5772–5788 66. Zhou D, Lin M, Chen Z, Sun H, Zhang H, Sun H, Yang B (2011) Simple synthesis of highly luminescent water-soluble CdTe quantum dots with controllable surface functionality. Chem Mater 23:4857–4862 67. Heuer-Jungemann A, Feliu N, Bakaimi I, Hamaly M, Alkilany A, Chakraborty I, Masood A, Casula MF, Kostopoulou A, Oh E, Susumu K, Stewart MH, Medintz IL, Stratakis E, Parak WJ, Kanaras AG (2019) The role of ligands in the chemical synthesis and applications of inorganic nanoparticles. Chem Rev 119:4819–4880 68. Sperling RA, Parak WJ (2010) Surface modification, functionalization and bioconjugation of colloidal inorganic nanoparticles. Philos Trans R Soc A 368:1333–1383 69. Susumu K, Uyeda HT, Medintz IL, Pons T, Delehanty JB, Mattoussi H (2007) Enhancing the stability and biological functionalities of quantum dots via compact multifunctional ligands. J Am Chem Soc 129:13987–13996 70. Zhou J, Liu Y, Tang J, Tang W (2017) Surface ligands engineering of semiconductor quantum dots for chemosensory and biological applications. Mater Today 20:360–376 71. Erathodiyil N, Ying JY (2011) Functionalization of inorganic nanoparticles for bioimaging applications. Acc Chem Res 44:925–935 72. Yu WW, Chang E, Falkner JC, Zhang J, Al-Somali AM, Sayes CM, Johns J, Drezek R, Colvin VL (2007) Forming biocompatible and nonaggregated nanocrystals in water using amphiphilic polymers. J Am Chem Soc 129:2871–2879 73. Wu X, Liu H, Liu J, Haley KN, Treadway JA, Larson JP, Ge N, Peale F, Bruchez MP (2003) Immunofluorescent labeling of cancer marker Her2 and other cellular targets with semiconductor quantum dots. Nat Biotechnol 21:41–46

Reprinted from the journal

171

13



Topics in Current Chemistry (2020) 378:35

74. Dubertret B, Skourides P, Norris DJ, Noireaux V, Brivanlou AH, Libchaber A (2002) In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 298:1759–1762 75. Park J, Lee J, Kwag J, Baek Y, Kim B, Yoon CJ, Bok S, Cho S-H, Kim KH, Ahn GO, Kim S (2015) Quantum dots in an amphiphilic polyethyleneimine derivative platform for cellular labeling, targeting, gene delivery, and ratiometric oxygen sensing. ACS Nano 9:6511–6521 76. Zhou J, Yang Y, Zhang C-Y (2015) Toward biocompatible semiconductor quantum dots: from biosynthesis and bioconjugation to biomedical application. Chem Rev 115:11669–11717 77. Qu W, Zuo W, Li N, Hou Y, Song Z, Gou G, Yang J (2017) Design of multifunctional liposomequantum dot hybrid nanocarriers and their biomedical application. J Drug Target 25:661–672 78. Green NM (1963) The use of ­[14C]biotin for kinetic studies and for assay. Biochem J 89:585–591 79. Biju V (2014) Chemical modifications and bioconjugate reactions of nanomaterials for sensing, imaging, drug delivery and therapy. Chem Soc Rev 43:737–962 80. Russ Algar W (2017) A brief introduction to traditional bioconjugate chemistry, Chap. 1. In: Algar WR, Dawson PE, Medintz IL (eds) Chemoselective and bioorthogonal ligation reactions: concepts and applications, vol 1. Wiley‐VCH Verlag GmbH. & Co. KGaA. In the other chapters of volume 1 the reactions mentioned in chapter 1 are dealt in detail 81. Bioconjugation technical handbook. Reagents for crosslinking, immobilization, modification, biotinylation, and fluorescent labeling of proteins and peptides. Thermo Fisher Scientific. https​://www. therm​ofish​er.com/es/es/home/globa​l/forms​/life-scien​ce/bioco​njuga​tion-techn​ical-handb​ook-downl​ oad.html. Accessed 11 Mar 2020. 82. Wagner AM, Knipe JM, Orive G, Peppas NA (2019) Quantum dots in biomedical applications. Acta Biomater 94:44–63 83. Lišková M, Voráčová I, Klepárník K, Hezinová V, Přikryl J, Foret F (2011) Conjugation reactions in the preparations of quantum dot-based immunoluminescent probes for analysis of proteins by capillary electrophoresis. Anal Bioanal Chem 400:369–379 84. Algar WR, Prasuhn DE, Stewart MH, Jennings TL, Blanco-Canosa JB, Dawson PE, Medintz IL (2011) The controlled display of biomolecules on nanoparticles: a challenge suited to bioorthogonal chemistry. Bioconjugate Chem 22:825–858 85. Massey M, Algar WR (2017) Nanoparticle bioconjugates: materials that benefit from chemoselective and bioorthogonal ligation chemistries. In: Algar WR, Dawson PE, Medintz IL (eds) Chemoselective and bioorthogonal ligation reactions: concepts and applications, vol 1. Wiley-VCH Verlag GmbH & Co. KGaA, Wienheim, pp 568–586 86. Row RD, Prescher JA (2018) Constructing new bioorthogonal reagents and reactions. Acc Chem Res 51:1073–1081 87. Wen L, Qiu L, Wu Y, Hu X, Zhang X (2017) Aptamer-modified semiconductor quantum dots for biosensing applications. Sensors 17:1736 88. Zhang C, Ding C, Zhou G, Xue Q, Xian Y (2017) One-step synthesis of DNA functionalized cadmium-free quantum dots and its application in FRET-based protein sensing. Anal Chim Acta 957:63–69 89. Long Y, Zhang LF, Zhang Y, Zhang CY (2012) Single quantum dot based nanosensor for renin assay. Anal Chem 84:8846–8852 90. Sanvicens N, Pascual N, Fernández-Argüelles MT, Adrián J, Costa-Fernández JM, Sánchez-Baeza F, Sanz-Medel A, Marco MP (2011) Quantum dot-based array for sensitive detection of Escherichia coli. Anal Bioanal Chem 399:2755–2762 91. Wang Y, Gao D, Zhang P, Gong P, Chen C, Gao G, Cai L (2014) A near infrared fluorescence resonance energy transfer based aptamer biosensor for insulin detection in human plasma. Chem Commun 50:811–813 92. Chen L, Tse WH, Chen Y, McDonald MW, Melling J, Zhang J (2017) Nanostructured biosensor for detecting glucose in tear by applying fluorescence resonance energy transfer quenching mechanism. Biosens Bioelectron 91:393–399 93. Cayuela A, Soriano ML, Carrillo-Carrión C, Valcárcel M (2016) Semiconductor and carbon-based fluorescent nanodots: the need for consistency. Chem Commun 52:1311–1326 94. Resch-Genger U, Grabolle M, Cavaliere-Jaricot S, Nitschke R, Nann T (2008) Quantum dots versus organic dyes as fluorescent labels. Nat Methods 5:763–775 95. Wu S, Liu L, Li G, Jing F, Mao H, Jin Q, Zhai W, Zhang H, Zhao J, Jia C (2016) Multiplexed detection of lung cancer biomarkers based on quantum dots and microbeads. Talanta 156–157:48–54

13

172

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35 96. Zhang W, Hubbard A, Brunhoeber P, Wang Y, Tang L (2013) Automated multiplexing quantum dots in  situ hybridization assay for simultaneous detection of ERG and PTEN gene status in prostate cancer. J Mol Diagn 15:754–764 97. Han M, Gao X, Su JZ, Nie S (2001) Quantum-dot-tagged microbeads for multiplexed optical coding of biomolecules. Nat Biotechnol 19:631–635 98. Kim J, Biondi MJ, Feld JJ, Chan WCW (2016) Clinical validation of quantum dot barcode diagnostic technology. ACS Nano 10:4742–4753 99. Zhao WW, Wang J, Zhu YC, Xu JJ, Chen HY (2015) Quantum dots: electrochemiluminescent and photoelectrochemical bioanalysis. Anal Chem 87:9520–9531 100. Campuzano S, Yáñez-Sedeño P, Pingarrón JM (2019) Nanoparticles for nucleic-acid-based biosensing: opportunities, challenges, and prospects. Anal Bioanal Chem 411:1791–1806 101. Zheng Y, Wang X, He S, Gao Z, Di Y, Lu K, Li K, Wang J (2019) Aptamer-DNA concatamer-quantum dots based electrochemical biosensing strategy for green and ultrasensitive detection of tumor cells via mercury free anodic stripping voltammetry. Biosens Bioelectron 126:261–268 102. Li CC, Hu J, Lu M, Zhang CY (2018) Quantum dot-based electrochemical biosensor for stripping voltammetric detection of telomerase at the single-cell level. Biosens Bioelectron 122:51–57 103. Liu Y, Zhu L, Kong J, Yang P, Liu B (2013) A quantum dots-based electrochemical assay towards the sensitive detection of tumor cells. Electrochem Comm 33:59–62 104. Martín-Yerga D, González-García MB, Costa-García A (2014) Electrochemical immunosensor for anti-tissue transglutaminase antibodies based on the in situ detection of quantum dots. Talanta 130:598–602 105. Kong FY, Xu BY, Xu JJ, Chen HY (2013) Simultaneous electrochemical immunoassay using CdS/ DNA and PbS/DNA nanochains as labels. Biosens Bioelectron 39:177–182 106. Christodouleas DC, Kaur B, Chorti P (2018) From point-of-care testing to eHealth diagnostic devices (eDiagnostics). ACS Cent Sci 4:1600–1616 107. Tavares AJ, Noor MO, Vannoy CH, Algar WR, Krull UJ (2012) On-chip transduction of nucleic acid hybridization using spatial profiles of immobilized quantum dots and fluorescence resonance energy transfer. Anal Chem 84:312–319 108. Noor MO, Tavares AJ, Krull UJ (2013) On-chip multiplexed solid-phase nucleic acid hybridization assay using spatial profiles of immobilized quantum dots and fluorescence resonance energy transfer. Anal Chim Acta 788:148–157 109. Kim C, Hoffmann G, Searson PC (2017) Integrated magnetic bead–quantum dot immunoassay for malaria detection. ACS Sens 2:766–772 110. Medina-Sánchez M, Miserere S, Morales-Narváez E, Merkoçi A (2014) On-chip magneto-immunoassay for Alzheimer’s biomarker electrochemical detection by using quantum dots as labels. Biosens Bioelectron 54:279–284 111. Krejcova L, Nejdl L, Merlos-Rodrigo MA, Zurek M, Matousek M, Hynek D, Zitka O, Kopel P, Adam V, Kizek R (2014) 3D printed chip for electrochemical detection of influenza virus labeled with CdS quantum dots. Biosens Bioelectron 54:421–427 112. Zhou F, Lu M, Wang W, Bian ZP, Zhang JR, Zhu JJ (2010) Electrochemical immunosensor for simultaneous detection of dual cardiac markers based on a poly(dimethylsiloxane)-gold nanoparticles composite microfluidic chip: a proof of principle. Clin Chem 56:1701–1707 113. Gong MM, Sinton D (2017) Turning the page: advancing paper-based microfluidics for broad diagnostic application. Chem Rev 117:8447–8480 114. Roda A, Michelini E, Zangheri M, Di Fusco M, Calabria D, Simoni P (2016) Smartphone-based biosensors: a critical review and perspectives. Trends Anal Chem 79:317–325 115. Bahadır EB, Sezgintürk MK (2016) Lateral flow assays: principles, designs and labels. Trends Anal Chem 82:286–306 116. Martinez AW, Phillips ST, Butte M, Whitesides GM (2007) Patterned paper as a platform for inexpensive, low-volume, portable bioassays. Angew Chem 119:1340–1342 117. Urusov AE, Zherdev AV, Dzantiev BB (2019) Towards lateral flow quantitative assays: detection approaches. Biosensors 9:89 118. Fu LM, Wang YN (2018) Detection methods and applications of microfluidic paper-based analytical devices. Trends Anal Chem 107:196–211 119. Sapountzi EA, Tragoulias SS, Kalogianni DP, Ioannou PC, Christopoulos TK (2015) Lateral flow devices for nucleic acid analysis exploiting quantum dots as reporters. Anal Chim Acta 864:48–54 Reprinted from the journal

173

13

Topics in Current Chemistry (2020) 378:35 120. Savin M, Mihailescu C-M, Matei I, Stan D, Moldovan CA, Ion M, Baciu I (2018) A quantum dotbased lateral flow immunoassay for the sensitive detection of human heart fatty acid binding protein (hFABP) in human serum. Talanta 178:910–915 121. Borse V, Srivastava R (2019) Fluorescence lateral flow immunoassay based point-of-care nanodiagnostics for orthopedic implant-associated infection. Sens Actuators B Chem 280:24–33 122. Qina W, Wang K, Xiao K, Hou Y, Lu W, Xu H, Wo Y, Feng S, Cui D (2017) Carcinoembryonic antigen detection with “Handing”-controlled fluorescence spectroscopy using a color matrix for point-ofcare applications. Biosens Bioelectron 90:508–515 123. Liu J, Ji D, Meng H, Zhang L, Wang J, Huang Z, Chen J, Li J, Li Z (2018) A portable fluorescence biosensor for rapid and sensitive glutathione detection by using quantum dots-based lateral flow test strip. Sens Actuators B Chem 262:486–492 124. Yan X, Wang K, Lu W, Qin W, Cui D, He J (2016) CdSe/ZnS quantum dot-labeled lateral flow strips for rapid and quantitative detection of gastric cancer carbohydrate antigen 72-4. Nanoscale Res Lett 11:138 125. Deng X, Wang C, Gao Y, Li J, Wen W, Zhang X, Wang S (2018) Applying strand displacement amplification to quantum dots-based fluorescent lateral flow assay strips for HIV-DNA detection. Biosens Bioelectron 105:211–217 126. Shah KG, Singh V, Kauffman P, Abe K, Yager P (2018) Mobile phone ratiometric imaging enables highly sensitive fluorescence lateral flow immunoassays without external optical filters. Anal Chem 90:6967–6974 127. Wang C, Hou F, Man Y (2015) Simultaneous quantitative detection of multiple tumor markers with a rapid and sensitive multicolor quantum dots based immunochromatographic test strip. Biosens Bioelectron 68:156–162 128. Wu F, Yuan H, Zhou C, Mao M, Liu Q, Shen H, Cen Y, Qin Z, Ma L, Li LS (2016) Multiplexed detection of influenza A virus subtype H5 and H9 via quantum dot-based immunoassay. Biosens Bioelectron 77:464–470 129. Rong Z, Wang Q, Sun N, Jia X, Wang K, Xiao R, Wang S (2019) Smartphone-based fluorescent lateral flow immunoassay platform for highly sensitive point-of-care detection of Zika virus nonstructural protein 1. Anal Chim Acta 1055:140–147 130. Li X, Li W, Yang Q, Gong X, Guo W, Dong C, Liu J, Xuan L, Chang J (2014) Rapid and quantitative detection of prostate specific antigen with a quantum dot nanobeads-based immunochromatography test strip. Appl Mater Interfaces 6:6406–6414 131. Petryayeva E, Algar WR (2015) Single-step bioassays in serum and whole blood with a smartphone, quantum dots and paper in-PDMS chip. Analyst 140:4037–4045 132. Das P, Krull UJ (2017) Detection of a cancer biomarker protein on modified cellulose paper by fluorescence using aptamer-linked quantum dots. Analyst 142:3132–3135 133. Noor MO, Hrovat D, Moazami-Goudarzi M, Espie GS, Krull UJ (2015) Ratiometric fluorescence transduction by hybridization after isothermal amplification for determination of zeptomole quantities of oligonucleotide biomarkers with a paper-based platform and camera-based detection. Anal Chim Acta 885:156–165 134. Lin YY, Wang J, Liu G, Wu H, Wai CM, Lin Y (2008) A nanoparticle label/immunochromatographic electrochemical biosensor for rapid and sensitive detection of prostate-specific antigen. Biosens Bioelectron 23(11):1659–1665 135. Kokkinos CT, Giokas DL, Economou AS, Petrou PS, Kakabakos SE (2018) Paper-based microfluidic device with integrated sputtered electrodes for stripping voltammetric determination of DNA via quantum dot labeling. Anal Chem 90:1092–1097 136. Cincotto FH, Fava EL, Moraes FC, Fatibello-Filho O, Faria RC (2019) A new disposable microfluidic electrochemical paper-based device for the simultaneous determination of clinical biomarkers. Talanta 195:62–68 137. Yang M, Liu Y, Jiang X (2019) Barcoded point-of-care bioassays. Chem Soc Rev 48:850–884 138. Gao Y, Lam AWY, Chan WCW (2013) Automating quantum dot barcode assays using microfluidics and magnetism for the development of a point-of-care device. ACS Appl Mater Interfaces 5:2853–2860 139. Ming K, Kim J, Biondi MJ, Syed A, Chen K, Lam A, Ostrowski M, Rebbapragada A, Feld JJ, Chan WCW (2015) Integrated quantum dot barcode smartphone optical device for wireless multiplexed diagnosis of infected patients. ACS Nano 9:3060–3074 140. Wegner KD, Hildebrandt N (2015) Quantum dots: bright and versatile in vitro and in vivo fluorescence imaging biosensors. Chem Soc Rev 44:4792–4834

13

174

Reprinted from the journal

Topics in Current Chemistry (2020) 378:35 141. Zhao M-X, Zeng E-Z (2015) Application of functional quantum dot nanoparticles as fluorescence probes in cell labeling and tumor diagnostic imaging. Nanoscale Res Lett 10:171 142. Liu YS, Sun Y, Vernier PT, Liang CH, Chong SYC, Gundersen MA (2007) pH-sensitive photoluminescence of CdSe/ZnSe/ZnS quantum dots in human ovarian cancer cells. J Phys Chem C 111:2872–2878 143. Ding H, Yong K-T, Law W-C, Roy I, Hu R, Wu F, Zhao W, Huang K, Erogbogbo F, Bergeya EJ, Prasad PN (2011) Non-invasive tumor detection in small animals using novel functional Pluronic nanomicelles conjugated with anti-mesothelin antibody. Nanoscale 3:1813–1822 144. Zhang LW, Bäumer W, Monteiro-Riviere NA (2011) Cellular uptake mechanisms and toxicity of quantum dots in dendritic cells. Nanomedicine 6:777–791 145. Xiao Y, Forry SP, Gao X, Holbrook RD, Telford WG, Tona A (2010) Dynamics and mechanisms of quantum dot nanoparticle cellular uptake. J Nanobiotechnol 8:13 146. Igor L, Medintz HTU, Goldman ER, Mattoussi H (2005) Quantum dot bioconjugates for imaging, labelling and sensing. Nat Mater 4:435–446 147. Rosenthal SJ, Chang JC, Kovtun O, McBride JR, Tomlinson ID (2011) Biocompatible quantum dots for biological applications. Chem Biol 18:10–24 148. Ni X, Castanares M, Mukherjee A, Lupold SE (2011) Nucleic acid aptamers: clinical applications and promising new horizons. Curr Med Chem 18:4206–4214 149. Tang J, Huang N, Zhang X, Zhou T, Tan Y, Pi J, Pi L, Cheng S, Zheng H, Cheng Y (2017) Aptamerconjugated PEGylated quantum dots targeting epidermal growth factor receptor variant III for fluorescence imaging of glioma. Int J Nanomed 12:3899–3911 150. McHugh KJ, Jing L, Behrens AM, Jayawardena S, Tang W, Gao M, Langer R, Jaklenec A (2018) Biocompatible semiconductor quantum dots as cancer imaging agents. Adv Mater 30:e1706356 151. Aswathy RG, Yoshida Y, Maekawa T, Kumar DS (2010) Near-infrared quantum dots for deep tissue imaging. Anal Bioanal Chem 397:1417–1435 152. Allen PM, Liu W, Chauhan VP, Lee J, Ting AY, Fukumura D, Jain RK, Bawendi MG (2010) InAs(ZnCdS) quantum dots optimized for biological imaging in the near-infrared. J Am Chem Soc 132:470–471 153. Liu XY, Braun GB, Zhong HZ, Hall DJ, Han WL, Qin MD, Zhao CZ, Wang MN, She ZG, Cao CB, Sailor MJ, Stallcup WB, Ruoslahti E, Sugahara KN (2016) Tumor-targeted multimodal optical imaging with versatile cadmium-free quantum dots. Adv Funct Mater 26:267–276 154. Shen Y, Lifante J, Ximendes E, Santos HDA, Ruiz D, Juárez BH, Gutiérrez IZ, Vera VT, Retama JR, Rodríguez EM, Ortgies DH (2019) Perspectives for Ag2S NIR-II nanoparticles in biomedicine: from imaging to multifunctionality Nanoscale. Jaque D, Benayas A, del Rosal B 11:19251–19264. https​:// doi.org/10.1039/C9NR0​5733A​ 155. Wang Z, Ma Y, Yu X, Niu Q, Han Z, Wang H, Li T, Fu D, Achilefu S, Qian Z, Gu Y (2018) Targeting CXCR4–CXCL12 axis for visualizing, predicting, and inhibiting breast cancer metastasis with theranostic AMD3100–Ag2S quantum dot probe. Adv Funct Mater 28:1800732 156. Hu J, Xie M, Wen C-Y, Zhang Z-L, Xie H-Y, Liu A-A, Chen Y-Y, Zhou S-M, Pang DW (2011) A multicomponent recognition and separation system established via fluorescent, magnetic, dualencoded multifunctional bioprobes. Biomaterials 32:1177–1184 157. Xu H-L, Yang JJ, ZhuGe DL, Lin MT, Zhu QY, Jin BH, Tong MQ, Shen B-X, Xiao J, Zhao Y-Z (2018) Glioma-targeted delivery of a theranostic liposome integrated with quantum dots, superparamagnetic iron oxide, and cilengitide for dual-imaging guiding cancer surgery. Adv Healthc Mater 7:1701130 158. Deng Y, Xu A, Yu Y, Fu C, Liang G (2018) Biomedical applications of fluorescent and magnetic resonance imaging dual-modality probe. ChemBioChem 20:499 159. Yang Y, Lin L, Jing L, Yue X, Dai Z (2017) C ­ uInS2/ZnS quantum dots conjugating Gd(III) chelates for near-infrared fluorescence and magnetic resonance bimodal imaging. ACS Appl Mater Interfaces 9:23450–23457 160. Estelrich J, Sánchez-Martín MJ, Busquets MA (2015) Nanoparticles in magnetic resonance imaging: from simple to dual contrast agents. Int J Nanomed 10:1727–1741 161. Garcia-Cortes M, Ruiz Encinar J, Costa-Fernandez JM, Sanz-Medel A (2016) Highly sensitive nanoparticle-based immunoassays with elemental detection: application to prostate-specific antigen quantification. Biosens Bioelectron 85:128–134

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. Reprinted from the journal

175

13



Topics in Current Chemistry (2020) 378:35

Affiliations María Díaz‑González1 · Alfredo de la Escosura‑Muñiz1 · Maria Teresa Fernandez‑Argüelles2 · Francisco Javier García Alonso3 · Jose Manuel Costa‑Fernandez2  1

NanoBioAnalysis Group, Department of Physical and Analytical Chemistry, University of Oviedo, Julián Clavería 8, 33006 Oviedo, Spain

2

GEAB Research Group, Department of Physical and Analytical Chemistry, University of Oviedo, Avda. Julián Clavería 8, 33006 Oviedo, Spain

3

NanoBioAnalysis Group, Department of Organic and Inorganic Chemistry, University of Oviedo, Avda. Julián Clavería, 8, 33006 Oviedo, Spain



13

176

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15 https://doi.org/10.1007/s41061-019-0278-8 REVIEW

Carbon Nanotubes in Biomedicine Viviana Negri1   · Jesús Pacheco‑Torres2   · Daniel Calle3   · Pilar López‑Larrubia4  Received: 18 October 2019 / Accepted: 31 December 2019 / Published online: 14 January 2020 © Springer Nature Switzerland AG 2020

Abstract Nowadays, biomaterials have become a crucial element in numerous biomedical, preclinical, and clinical applications. The use of nanoparticles entails a great potential in these fields mainly because of the high ratio of surface atoms that modify the physicochemical properties and increases the chemical reactivity. Among them, carbon nanotubes (CNTs) have emerged as a powerful tool to improve biomedical approaches in the management of numerous diseases. CNTs have an excellent ability to penetrate cell membranes, and the sp2 hybridization of all carbons enables their functionalization with almost every biomolecule or compound, allowing them to target cells and deliver drugs under the appropriate environmental stimuli. Besides, in the new promising field of artificial biomaterial generation, nanotubes are studied as the load in nanocomposite materials, improving their mechanical and electrical properties, or even for direct use as scaffolds in body tissue manufacturing. Nevertheless, despite their beneficial contributions, some major concerns need to be solved to boost the clinical development of CNTs, including poor solubility in water, low biodegradability and dispersivity, and toxicity problems associated with CNTs’ interaction with biomolecules in tissues and organs, including the possible effects in the proteome and genome. This review performs a wide literature analysis to present the main and latest advances in the optimal design and characterization of carbon nanotubes with biomedical applications, and their capacities in different areas of preclinical research. Keywords  Carbon nanotubes · Biomedical research · Preclinical applications · Cancer · Neurodegeneration · Imaging · Theranostic compounds · Tissue engineering

Chapter 6 was originally published as Negri, V., Pacheco‑Torres, J., Calle, D. & López‑Larrubia, P. Topics in Current Chemistry (2020) 378: 15. https://doi.org/10.1007/s41061-019-0278-8. * Pilar López‑Larrubia [email protected] Extended author information available on the last page of the article Reprinted from the journal

177

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:15

Abbreviations AFM Atomic force microscopy AGP Angiopep-2 BBB Blood–brain barrier BLI Bioluminescence imaging BRB Berberine CA(s) Contrast agent(s) CNT(s) Carbon nanotube(s) DMF Dimethylformamide DNA Deoxyribonucleic acid DOX Doxorubicin EM Electron microscopy FTIR Fourier-transformed infrared spectroscopy Gd Gadolinium GNTs Gado-nanotubes HA Hyaluronic acid MRI Magnetic resonance imaging MWCNT(s) Multi-walled carbon nanotube(s) NGF Nerve growth factor NIR Near-infrared radiation NP(s) Nanoparticle(s) PEG Polyethylene glycol PET Positron emission tomography PLK1 Polo-like kinase 1 PTT Photothermal therapy RNA Ribonucleic acid SC Stem cells SDBS Sodium dodecyl benzene sulfonate SEM Scanning electron microscopy siRNA Small interfering ribonucleic acid SPECT Single-photon emission computed tomography STM Scanning tunneling microscopy SWCNT(s) Single-walled carbon nanotube(s) TEM Transmission electron microscopy XPS X-ray photoelectron spectroscopy

1 Introduction Nanotechnology is the understanding, design, and development of new materials with interesting characteristics at the nanometric scale where the effects of quantum mechanics are noticeable. These nanomaterials have at least one dimension between 1 and 100  nm, and adapt the shape of nanoparticles, nanotubes, or nanosurfaces. On the other side, nanobiotechnology refers to the intersection of nanotechnology and biology, combining the efficacy of biological materials and the rules and tools of basic sciences like physics and chemistry. It is also defined as the science that

13

178

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



investigates, beyond the molecular level, the design and development of highly ordered structured materials that offer specific responses when exposed to stimuli. The development of these nanomaterials has reached a wide use in numerous scientific and technical fields, including water treatment [1, 2], electrocatalysis [3, 4], catalysis [5–9], materials [10], and biomedical applications from cancer treatment to regenerative medicine [11–13]. In fact, these structures have become increasingly useful in two main areas of biomedicine: (1) nanomedicine, with noteworthy applications in imaging, biosensors, drug delivery systems, and photo thermal therapy; and (2) tissue and implants engineering either as scaffold-based nanomaterials or as components of biomedical devices. These nanocompounds can be subcategorized into four wide groups depending on their composition: carbon-based, inorganic-based, organic-based, and composite-based. Among the carbon-based materials, carbon nanotubes (CNTs)—found by Ijima in 1991—constitute a new allotrope of carbon [14] (Fig. 1) that deserves special interest due to their inherent features (surface, shape, and physical properties) that make them especially suitable for preclinical applications [15]. CNTs are tubular structures made of a layer of graphene rolled up into a cylinder [16]. These NPs are classified according to the number of wall sheets in their structure as single-walled carbon nanotubes (SWCNTs), consisting of a single graphene sheet with diameter that typically varies in the range of 0.4–40 nm, and multi-walled carbon nanotubes (MWCNTs), consisting of multiple sheets forming concentric cylinders with an interlayer distance of 0.35 nm, similar to the basal plane separation in graphite, with diameters from 2 to 100 nm (Fig. 2a). The nanotubes are usually closed at the ends with half-fullerene molecules shape, with pentagonal defects that form the tips. CNTs can be also be categorized into three types, depending on the rolling up of the sheets (Fig. 2b) as armchair, zig-zag, or chiral nanotubes (Fig. 2b) [17].

Fig. 1  Carbon allotropes: diamond, graphite, lonsdaleite, C60-fullerene, graphene, amorphous carbon, C540-fullerite, and single-walled carbon nanotube Reprinted from the journal

179

13



Topics in Current Chemistry (2020) 378:15

Fig. 2  Carbon nanotube classifications. a According to the number of sheets, there are single- (SWCNTs) and multi-wall carbon nanotubes (MWCNTs); b depending on the rolling up of the sheets, they can be armchair, zig-zag, or chiral

In CNTs, each carbon atom is bonded with sp2 hybridization, stronger than the sp3 bonds in diamond, which provide these compounds with exceptional mechanical [18–20], electrical [21–24], optical [25], and thermal properties [26]. Other unique features owing to their large surface, needle shape, and residual metal impurities content placed them among the most promising nanomaterials for potential applications that range from nanomedicine to nanoelectronics, including the production of quantum dots by introducing fullerenes [27] and nanowires by filling CNT with pure elements for molecular electronics [28]. Regarding the use of carbon nanotubes in biomedicine and preclinical research, and due to their ability to cross the cell membrane, these materials has been tested as transporters for different drugs [29], biomolecules as enzymes, deoxyribonucleic acid (DNA) and ribonucleic acid (RNA) [30], as well as to form ion transport channels [31]. CNTs have also been used as nanoreactors, taking advantage of their reduced inner size and the special electronic characteristics due to the CNT wall curvature [32]. Nevertheless, despite their wide range of biomedical applications, carbon nanotubes have two deficiencies to solve: their inherent toxicity, due to the metal catalyst residue remaining from the synthesis process, and their low solubility in water. On these grounds, we have focused this review on the presentation of the main functionalization processes, characterization methodologies, and biomedical applications of CNTs, identifying also the most relevant strategies to overcome the inherent problems that currently hinder their preclinical use.

2 Functionalization of Carbon Nanotubes Carbon nanotubes are produced using several synthetic methods [33], but most of them left metallic impurities in the sample and render a mixture of CNTs with different diameters, lengths, and chirality properties. Additional problems are the

13

180

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



tendency of nanotubes to agglomerate uncontrollably due to the high surface energy, and the stabilization of the bundles by van der Waals forces and π–π electron interactions among them. This phenomenon entails weak dispersibility and makes CNTs insoluble in most biocompatible solvents, highly limiting their biomedical applications. Functionalization solves many of these problems by modifying the CNT surface properties. Through functionalization, CNTs can be purified achieving high homogeneity, decreased toxicity, increased dispersibility, and solubility [34]. It also allows specific CNT decoration for different purposes and applications. This review presents the main functionalization pathways, with specific examples based on preparing nanotubes for preclinical research. There are multiple strategies for functionalization of CNTs with applications in biomedicine [35]. Traditionally, they have been categorized according to the functionalization approaches into covalent and non-covalent, but a higher hierarchy factor can be introduced: the location of the functionalization. Endohedral functionalization accounts for changes in the inner face of the CNTs, whereas exohedral functionalization aims to add new functionalities in the outer face. Endohedral functionalization does not involve the formation of bonds between CNT and functional groups, but filling the inner part of the CNT with host material [36]. The first described endohedral functionalization was the filling with water through capillarity [37], but soon other solvents followed [38]. Hydrophobic molecules can also be successfully entrapped in the inner space of CNTs by simple incubation [39, 40]. Some other examples of endohedral functionalization by fullerene encapsulation were reported by Karousis et al. [41]. Nevertheless, this synthetic approach is limited due to the small diameter of the inner cavity of the CNTs, greatly restricting the size of the encapsulated elements. Exohedral functionalization, on the other hand, aims to decorate the CNTs in their outer face. In principle, it has no size limitations and can potentially attach almost any type of functional group. The covalent strategies imply the formation of new chemical bonds altering the original CNT structure, and the non-covalent ones are based on the interaction through π–π stacking forces and/or van der Waals preserving the CNT skeleton intact. We present here the different types of functionalization categorized for clarity purposes, but it is important to notice that most of the synthetic procedures employed nowadays use a combination of several of the methodologies described here. 2.1 Covalent Functionalization Covalent functionalization implies the formation of new chemical bonds, altering the original CNT structure. This reaction will change the carbon hybridization from sp2 to sp3, causing the loss of the π-conjugation system of the graphene layer [42] responsible for most of the optical, electrical, and thermal properties of CNTs [43]. Despite this drawback, covalent functionalization presents many characteristics that make this approach highly attractive: it provides strong and very stable attachment of functional groups, entails higher selectivity, is more robust and better controlled

Reprinted from the journal

181

13



Topics in Current Chemistry (2020) 378:15

than non-covalent functionalization, can be done in organic solvent or even without solvent, and offers a huge plethora of functional groups that can be used [44]. All these benefits highly assist the potential use of carbon nanotubes in biomedicine. Briefly, covalent functionalization can be classified depending on whether the modifications are performed at the sidewalls or in defect sites. 2.1.1 Side Wall Functionalization The sidewalls of carbon nanotubes are considered to be very inert, so their direct functionalization will only occur if a highly reactive agent is used [45]. Singh et al. [46] widely describe the main side-wall derivatization strategies in SWCNTs, including halogenation, arylation, nucleophilic addition, radical addition, cycloadditions, and carboxyl chemistry reactions. Here we present the most commonly used strategies in preparing CNTs in order to increase their solubility and dispersivity [34], decrease the inherent toxicity [47], and improve their biocompatibility with biomedical purposes. 2.1.1.1  Halogenation  Fluorination was first used to overcome the lack of CNT reactivity using elemental fluorine at temperatures between 25 and 600 °C [48, 49]. These new C-F bonds are weaker than those in alkyl fluorides [50] and can be employed for further functionalization [51], replacing fluorine with amines [52], alcohols [53], or alkyl groups using Grignard [54] or organolytic reagents [55] (Scheme 1). Besides fluorination, chlorination and bromination of CNTs can also be achieved using electrolysis [56]. 2.1.1.2  Electrophilic and Nucleophilic Additions  Electrophilic addition of alkyl halides results in the formation of alkyl and hydroxyl groups, whereas nucleophilic addition of amine-based nucleophiles leads to amino-functionalized CNTs (Scheme 2). As an example of electrophilic reaction, Friedel–Crafts acylation between MWCNT

Scheme 1  Fluorination and further functionalization of carbon nanotubes

13

182

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

and p-aminobenzoic acid in the presence of polyphosphoric acid renders MWCNTNH2 (Scheme 2a), which can be used for further functionalization with poly(l-lactide) polymer [57] or with collagen using glutamic acid as a crosslinker [58]. As an example of nucleophilic addition, SWCNT can be alkylated by the treatment with t-butyllithium and subsequent reoxidation of the intermediates to obtain neutral CNTs decorated with tert-butyl moieties [59]. 2.1.1.3  Radical Additions  Originally, this approach was developed using substituted aryl diazonium salts electrochemically reduced in organic media (Scheme 2c) [60, 61]. The electron transfer between the CNT and the aryl diazonium salt triggered the formation of the aryl radicals. Posterior developments allowed performing this reaction in water [62] and to generate highly functionalized carbon nanotubes using micelle-coated CNTs [63, 64]. In situ generation of the diazonium salt provided functionalized well-dispersed nanotubes in organic solvents [65, 66], in aqueous solutions [67] and in solvent-free conditions [68]. Finally, electrochemistry allows both the reductive and the oxidative attachment of substituted phenyl groups [69, 70].

Scheme 2  Surface derivatization of CNT: a Friedel–Crafts acylation; b electrophilic aromatic addition; c radical additions using diazonium salts reaction; d ozonolysis Reprinted from the journal

183

13



Topics in Current Chemistry (2020) 378:15

2.1.1.4  Cycloaddition  Cycloadditions are another important way of functionalizing the walls of nanotubes to tune their biocompatibility and biodegradability, both crucial characteristics to perform in vivo studies. Delgado and colleagues [71] described for the first time the [4 + 2] Diels–Alder reaction of o-quinodimethane assisted by microwaves on SWCNT surface (Scheme  3). More recently, anhydride-functionalized CNTs were produced using a cascade of Diels–Alder cycloaddition reactions employing 1,3-butadiene generated from 3-sulfolene in the presence of atmospheric oxygen [72]. A very versatile methodology uses 1,3-dipolar cycloadditions between the nanotube and azomethine ylides generated in situ by condensation of α-amino acids and aldehydes [73, 74]. This approach yields bi-substituted pyrrolidines and highly functionalized CNTs (Scheme 4) [75, 76]. Using this strategy, Calcio Guadino et al. [77] obtained multi-decorated SWCNTs in a single step. Single microwave-assisted grafting reaction labels nanotubes’ surface with amino acidic β-cyclodextrin derivative and the (1,4,7,10-tetraazacyclododecane-N,N’,N’’,N’’’-tetraacetic acid monoamide) moiety with a 1:1 ratio. 2.1.2 Defect Functionalization Defect functionalization makes use of the structural weaknesses of CNTs to create new bonds. Although nanotubes are highly unreactive, all the preparation methods to obtain them leave some defects in their structure, both in the lateral walls and in the tips. Among the lateral walls, these defects include dipoles of heptagon–pentagon pairs in the hexagonal network called Stone–Wales defects, sp3-hybridized defects, and vacancies in the sp2 network [45]. The nanotube’s ends are usually closed with fullerene, presenting mixed pentagonal–hexagonal structures more reactive than the pristine lateral walls [45]. Defect functionalization employs these intrinsic defects or generates new ones on the nanotube structure, normally by aggressive oxidative processes either in liquid or gas phase, or by electrochemical oxidation [78]. This approach has the advantage of generating more functional groups, but also implies higher structural damage. Using less aggressive oxidative approaches, the structural damage is minimized preserving CNT’s properties [79]. Defect functionalization can be classified according to the nature of the chemical transformation occurring at the defect sites, like oxidation, amidation, thiolation, etc. These attached functional groups are normally used as the starting point for further derivatization [80]. Oxidation and carboxyl-based coupling constitutes one of the most relevant strategies for CNTs functionalization, including the formation of esters [81, 82], amides [83], and ammonium carboxylate salts [84]. This modification is achieved in a twostep process. The first one often involves the treatment of commercially available carbon nanotubes with a mineral acid [85] such as nitric acid [86] or a sulfonitric mixture [87]. This step shortens CNTs, narrowing length distribution [88], and reduces the metal concentration left from their synthesis, decreasing toxicity and improving the preclinical possibilities. Nanotube defects will suffer the effects of the oxidant creating carboxylic and other oxygen-bearing groups [85]. The second step occurs via carboxylic acid formation, and can be carried out using two methodologies: (1) employing in situ acid chloride formation with thionyl chloride and

13

184

Reprinted from the journal

Scheme 3  [4 + 2] Diels–Alder reaction of o-quinodimethane assisted by microwaves on SWCNT surface

Topics in Current Chemistry (2020) 378:15

Reprinted from the journal

185

13



Topics in Current Chemistry (2020) 378:15

Scheme 4  1,3-Dipolar cycloaddition reaction between carbon nanotubes’ surface and azomethine ylide generated in situ

subsequent reaction with a primary amine or alcohol (Scheme  5a) [87]; (2) using activating agents, such as 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC) and N-hydroxysulfosuccinimide sodium salt (NHS), and subsequent reaction with a primary amine (Scheme 5b) [89]. 2.2 Non‑covalent Functionalization Non-covalent functionalization does not imply the formation of new chemical bonds. The linkage with the nanotube is achieved through van der Waals forces, π–π interactions, hydrogen bonds, and/or electrostatic interactions, preserving the CNT’s structure intact. As a result, non-covalent functionalization ensures the sp2-hybridized six-membered ring network and the extended π-conjugation, maintaining the physical, electric, thermal, and optical properties of CNTs. However, this comes at the cost of a weaker anchoring of the functional groups that can be detached by changes in environmental conditions, as modifications in pH, temperature, or solvent. This weaker bonding could be transformed into an advantage, making this functionalization reversible, something especially useful for therapeutic and

Scheme 5  Functionalization of carbon nanotubes through carboxylate via acid chloride (a) and EDCNHS (b)

13

186

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

theranostic purposes in preclinical research. Non-covalent functionalization can be categorized according to the derivatization process like: small molecule adsorption or π–π stacking attachments, use of surfactants, and interaction with biomolecules. 2.2.1 Small Molecule Adsorption or π–π Stacking Aromatic derivatives, such as pyrene, porphyrins, and aromatic macrocycles, can functionalize the sidewalls of CNT through π–π stacking interactions [88] or donor–acceptor systems [90]. 2.2.1.1  Pyrene Derivatives  Pyrene is the smallest peri-fused planar hydrocarbon. It is highly symmetrical, having 16π electron and aromatic character despite not following Hückel’s 4n + 2 rule [91]. Pyrene-containing molecules have been extensively used as bridges to functionalize nanotubes’ surface with different biomolecules, polymers, dendrimers, and DNA derivatives, among others [92]. Zhu and coworkers anchored CNTs onto oxide surfaces using bifunctional molecules, with succinimidyl ester and pyrene groups as a bridge [93]. Chen et  al. employed 1-pyrenebutanoic acid, succinimidyl ester to functionalize carbon nanotubes with proteins [90]. Calle et al. used 1-amine pyrene to obtain homogenous π–π stacking adducts with MWCNT [88]. Once the pyrenyl group is anchored to CNTs, they can be further decorated with ester groups that are highly reactive to primary and secondary amines ubiquitous on the surface of proteins. 2.2.1.2  Porphyrin Derivatives Porphyrin is a π-conjugated system in a planar arrangement with a total of 18 π electrons and aromatic character. Using this structure, Zhang and coworkers [94] prepared polymeric porphyrin-functionalized CNTs by the condensation of terephthaldehyde and pyrrole in the presence of carbon nanotubes. Subsequent metalation with ­Ru3(CO)12 incorporated ruthenium to the composite, which showed excellent catalytic performance toward hydrogenation of biomassrelated compounds. Derivatized porphyrins can also be employed [95]. 2.2.2 Surfactants Amphiphilic molecules interact with carbon nanotubes in aqueous media through their hydrophobic parts while their hydrophilic ends face outwards [96]. Surfactants have been widely used to suspend CNTs in aqueous solutions, increasing their dispersibility [97] and reducing their cytotoxicity [98]. Examples of this type of molecule include deoxycholic acid, sodium dodecyl sulfonate, sodium dodecylbenzene sulfonate, and sodium dodecyl sulfate, among others [96, 99, 100]. This procedure is very effective in dispersing carbon nanotubes, allowing studying CNTs individual properties, and can be used as the starting point for further functionalization. Furthermore, if the hydrophobic tails of the surfactant present aromatic moieties, particularly strong π–π stacking interaction will be established with the CNT, as with sodium dodecyl benzene sulfonate (SDBS) [101].

Reprinted from the journal

187

13



Topics in Current Chemistry (2020) 378:15

Some examples of this approach include the work of Niezabitowska et al. [102] that employed sodium dodecyl sulfate as anionic surfactant to disperse carbon nanotubes and polycaprolactone. Negri et  al. [103] used SWCNT oxidized with 2% SDBS suspensions to generate nanotubular paramagnetic probes, with potential utility as contrast agents (CAs) in magnetic resonance imaging (MRI). Cerpa et al. [104] used a SDBS solution to disperse single-walled carbon nanotubes to obtain anisotropic relaxation probes for diffusion MRI studies. Bharti et al. [105] compared the role of a cationic an anionic surfactants in the synthesis and resulting properties of MWCNTs decorated with bi-metallic Pt–Pd nanoparticles. Yasujima [106] used Triton X-100® to disperse CNTs in a bioanode preparation process in a multienzyme immobilized carbon-felt electrode. Martinez-Paz et  al. [107] employed 0.015% Pluronic F68 culture medium solution to obtain a homogenous suspension of oxidized MWCNTs in order to determine the possible toxic effects in invertebrate Chironomus riparius caused by CNTs environmental dispersion. 2.2.3 Biomolecules The non-covalent functionalization by attaching biomolecules to the CNTs surfaces is attracting great attention in biomedical research because of their promising preclinical possibilities. With this aim, different biological molecules and macromolecules can be bound to the nanotubes, like polypeptides, DNA bases, DNA oligonucleotides, amino acids, phospholipids, etc. 2.2.3.1  Proteins  Carbon nanotubes can interact directly with proteins through π–π stacking of their aromatic residues (Trp, Phe, and Tyr) enhancing their absorptivity and biocompatibility [108, 109] and so increasing the possibilities of being used in preclinical evaluations. A different approach uses a bridge to anchor proteins to CNTs, either through covalent modifications of nanotubes [110], or by a non-covalent method employing pyrene derivatives [90]. 2.2.3.2  DNA Derivatives  DNA can bind to carbon nanotubes, forming helices around them [111] or can form non-covalent conjugates through the π–π staking with the aromatic bases [30]. DNA-functionalized CNTs can be used as biological transporters and also as biosensors [111, 112]. 2.2.3.3  Phospholipids  Due to their amphiphilic nature, phospholipids can be used as surfactants to solubilize CNTs. Lysophospholipids, or single-chained phospholipids, are very efficient in this task [113]. In this case, the lipid part wraps the nanotubes as striations, whereas the hydrophilic part provides CNT solubility and biocompatibility. Using this approach, phospholipid-polyethylene glycol (PEG) is employed to functionalize SWCNT for a range of different biomedical applications [114] (Fig. 3). In a different approach, lipid bilayers can encapsulate CNTs, creating a model to study different biological process occurring at the cell membranes [115].

13

188

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

Fig. 3  Functionalization protocol of SWCNT with PEG (steps 1–5), and further conjugation of targeting ligands to SWNTs (A and B), radiolabeling of SWNT (C), siRNA conjugation (D), and doxorubicin loading onto functionalized SWNTs (E) (with permission from [114])

3 Characterization of Carbon Nanotubes As stated before, all the production methods for carbon nanotubes generate nonhomogeneous material, varying in diameter, length, chirality, purity, etc. Thus, adequate CNTs characterization is a fundamental step previous to the use of these NPs in biomedicine. Parameters to be studied include thermal stability, homogeneity, conductivity, optical properties, and identification of the functional groups in case of covalent modifications. Characterization is also key in determining the metal traces from the synthesis, estimating carbonaceous impurities and studying structural defects in the sidewalls and tips. The main techniques for CNT characterization include photoluminescence spectroscopy, X-ray photoelectron spectroscopy, electron microscopy, scanning tunneling microscopy, X-ray diffraction, neutron diffraction, Raman spectroscopy, thermal analysis, absorption spectroscopy, and infrared spectroscopy, among others [116]. It is important to bear in mind that despite the wide range of techniques available, multiple characterization techniques must be used to obtain a complete description of a carbon

Reprinted from the journal

189

13



Topics in Current Chemistry (2020) 378:15

nanotube sample. Also, measurements are highly dependent on sample preparation and specific protocol details, and most of the times a reference is needed [117]. As stated previously, this article does not pretend to be an extensive review on the characterization techniques used to study nanotubes, but we just would like to give a brief overview of the main techniques used for characterization of CNTs with biomedical applications. For a more detailed review, interested readers could go to Refs. [117–119]. 3.1 Raman Spectroscopy Raman spectroscopy is one of the most powerful and used techniques for carbon nanotube characterization [120]. It is fast, does not need sample preparation, and is non-destructive. For SWCNTs, Raman spectroscopy provides qualitative and quantitative information about diameter, purity, crystallinity, and electronic structure, allowing to distinguish between metallic and semiconducting CNTs [121]. Furthermore, it supports studying and bundle CNTs [122, 123]. The most characteristic bands of nanotubes in Raman spectra are: (1) A ­ 1g or “breathing mode”, related to the diameter of the tube, (2) D-line, assigned to residual ill-organized graphite, and (3) G-band, related to highly ordered CNT sidewalls. The ratio between D- and G-bands can provide quantitative information about sidewall damage and changes produced by functionalization [124]. 3.2 Electron Microscopy (EM) EM includes transmission electron microscopy (TEM) and scanning electron microscopy (SEM). These are essential tools for studying directly the local structure of CNT at the nanometer level. TEM allows determining lengths and outer and inner diameters (Fig.  4) [88]. It also gives a qualitative estimation of the metallic and carbonaceous impurities, which appear as dark dots in the images [104]. The main disadvantages of EM techniques include possible damage to the sample due to the high energy of the electronic beam, and the large impact of sample preparation and drying in the results. 3.3 Scanning Probe Microscopy Among the different techniques encompassed within the scanning techniques, the two most commonly used to characterize functionalized CNT are atomic force microscopy (AFM) and scanning tunneling microscopy (STM) [46, 125]. AFM can be used to evaluate the stiffness and strength of individual MWCNT as well to measure their size distribution [126]. STM can reveal the atomic structure and the electronic properties of individual SWCNT [127], and is also able to image functional groups attached to the nanotube [125].

13

190

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

Fig. 4  TEM images of oxidized MWCNTs at 200  kV (left), and aminopyrene–MWCNT π–π stacking adducts showing open ends and aminopyrene adsorption (right). The image on the right was acquired with a GRANDARM300cFEG microscope with corrective aberration in the objective lens

3.4 Fourier‑Transformed Infrared Spectroscopy (FTIR) FTIR provides information about the impurities remaining from the nanotube synthesis, the catalytic activity of CNTs [128], and the functional organic groups attached during functionalization [124]. As an example, Fig.  5 presents two

Fig. 5  FTIR spectra of pristine MWCNT (black), oxidized MWCNT (blue). Note peak at 1691 cm−1 due to C=O stretching and broad band at 3300 cm−1 O–H Reprinted from the journal

191

13



Topics in Current Chemistry (2020) 378:15

spectra of pristine and oxidized CNTs, where the band due to carbonyl stretching of carboxylic group is clearly visible at 1691  cm−1. The main disadvantages of this technique include their qualitative nature and that some modifications cannot be observed due to the weak infrared-associated signals. 3.5 X‑Ray Photoelectron Spectroscopy (XPS) XPS provides information about the chemical structure of CNT (except for hydrogen) and, most importantly, the structural modifications due to chemical functionalization [129]. This technique irradiates CNTs with X-rays and determines the binding energy of the ejected photoelectrons. As an example, Fig.  6 shows the spectra of different CNTs, where the relative increase in O1s’ peak confirms the presence of carboxylic groups on oxidized nanotubes. Some inconveniences of this methodology are the requirement of relatively large amounts of sample and that peak fitting can be ambiguous. 3.6 Thermogravimetric Analysis Thermogravimetric analysis measures changes in the mass of the CNTs over time as the temperature varies under a control atmosphere. This technique is used to assess the purity of the sample and the concentration of organic molecules attached to the nanotubes. Thermogravimetric analysis is based on the lower decomposition temperatures of the adsorbed molecules and amorphous carbon compared to pristine CNTs. When the assay is carried out under air, the sample is completely oxidized and the remaining constitute the metallic impurities [116]. The main disadvantages include the destruction of the sample, requirement of large amounts of sample, and that data interpretation is often subjective.

Fig. 6  XPS of MWCNT (black), oxidized MWCNT (blue) and US shortened MWCNT (green) (with permission from Calle et al. [88])

13

192

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



4 Biomedical Applications of Carbon Nanotubes As we anticipated in the previous sections, CNTs present some physical features very interesting for biomedical applications, including a large surface area [130], electrical [23] and thermal conductivity [131], and optimal mechanical properties [132]. Some of these applications are achievable through the CNT’s conjugation with different biomolecules or compounds like polymers, proteins, DNA and RNA, among others, as above described (Fig. 7). Besides, their needle shape permits them to cross biological membranes and access cells and tissues in a non-affordable way for most of the common drugs and compounds. Nevertheless, some challenges need to be overtaken to boost their potential in biomedicine. These include low biocompatibility [133], mainly due to poor water solubility, low dispersivity, and high toxicity [47]. Even so, numerous groups have reported different and very interesting applications for CNTs also in the preclinical and in the clinical settings. 4.1 Diagnostic applications Early diagnosis and proper monitoring of disease is vital for efficient treatment of illnesses. This encourages the development of improved methodologies that solve some of the inconveniences and handicaps of the current methods, including sensitivity and selectivity, spatial and temporal resolution, cost, etc. The use of nanoparticles in general, and carbon nanotubes in particular, offers a wide range of possibilities that could be key in achieving these improvements. 4.1.1 Biosensors Biosensors are devices incorporating biological elements with unique binding specificities towards target analytes. In this field, CNT’s characteristics make them very interesting to investigate as relevant constituents of electrochemical biosensors [135]. Nanotubes are well suited for transduction of electric signals generated

Fig. 7  Schematic representation of surface functionalization and loading of carbon nanotubes for biomedical applications (with permission from Fernandes et al. [134]) Reprinted from the journal

193

13



Topics in Current Chemistry (2020) 378:15

upon recognition of a target, so, numerous applications are reporting the design of nanotube-based biosensors to detect and monitor different pathologies [136, 137]. Recently, a three-dimensional network of carbon nanotubes on Si pillar substrate has been developed for the accurate detection of oral squamous cell carcinoma in clinical saliva samples [138]. In this work, Song and colleagues described the preparation of the sensor, the in  vitro characterization, and the clinical applicability. The results obtained with this new CNT network showed a good correlation with data obtained using the commercially available electrochemiluminescence detection system employed in the hospital. Also, a novel CNT-based biosensor has been used for ultrasensitive detection of hydrogen peroxide and glucose in human serum, with a great interest for basic research and disease diagnosis [139]. In this study, the authors constructed a ratiometric fluorescent nanosensor based on the peroxidaselike properties of a hierarchical cobalt/carbon nanotube hybrid nanocomplex. This system assay developed reaches a detection limit of H ­ 2O2 of 100 nM and a selective and sensitive detection of glucose as low as 150  nM. In a different approach, multivalent electrodes for glucose biosensing were constructed through multiple functionalization of CNTs [140]. Three different pyrene derivatives were simultaneously immobilized on the nanotube surface by π–π-stacking: adamantane-pyrene, biotinpyrene, and nitrilotriacetic. They were adsorbed on the nanotube sidewalls to allow the step-by-step immobilization, via supramolecular host–guest interactions, of β-cyclodextrin modified glucose oxidase, biotinylated glucose oxidase, and histidine modified glucose (Fig.  8). The calibration curves for the glucose responses were performed by amperometry and using glucose oxidase as an enzyme model for all immobilization steps. DNA detection is a very active research area holding great promise in the early detection of many diseases and pathological processes, and CNTs offer strong opportunities to achieve that [141]. An interesting study reported the synthesis of a multi-functional gold/iron-oxide nanoparticle-CNT as a virus DNA-sensing platform [142]. The authors prepared the sensor through a simple two-step method obtaining the hybrid nanostructure that exhibited excellent detection potential and DNA sensing performance for different diseases. Chen and collogues also reported the fabrication of a DNA nano-biosensor system containing carbon nanotubes to detect the presence of Mycobacterium tuberculosis rapidly and with a great sensitivity [143]. CNTs have also been experimentally and theoretically investigated as conducting channels in a chemiresistor for the electrochemical detection of doublestranded DNA [144]. More recently, the development of a CNT-based field-effect transistor for DNA hybridization detection was reported [145]. In this sensor, DNA can bind well to a suspended CNT, avoiding the adverse effects of a substrate on a sensing material, and reaching a detection limit up to 10 aM. 4.1.2 Imaging Carbon nanotubes can be powerful tools with diagnosis purposes not only as biosensors but also to be used in imaging technologies. Due to their excellent intrinsic properties, CNTs have been employed as CAs in photoimaging techniques and are good platforms to carry molecules that make them detectable with different imaging

13

194

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

Fig. 8  SWCNT multivalent glucose biosensor by a coating process with adamantane-pyrene, biotin-pyrene and nitrilotriacetic acid-pyrene, and further host–guest interaction with avidin and β-cyclodextrin

modalities, as positron emission tomography (PET) or magnetic resonance imaging, among others [146]. MRI is probably the most powerful and versatile of all imaging techniques used in the clinical routine, biomedical research, and preclinical studies. Among its advantages, MRI presents wide implementation, high cost efficiency, non-invasiveness, and it does not use ionizing radiation. Although this imaging technique offers a great contrast between pathological and healthy tissues, the use of contrast agents is often required. CNTs are highly explored as CA candidates, with great sensitivity and specificity, low dose and reduced side effects, with numerous in  vitro and in vivo studies reported in the literature [147]. The most direct approach is to use them as negative T2 CAs making use of the remaining metal catalyst employed in the synthesis of CNTs [148], but these structures can be also prepared as positive T1 contrast media either by addition or trapping of gadolinium (Gd) complex [149, 150]. In both situations, some issues have to be solved to boost their widespread use, like toxicity and dispersion capacity [34]. Despite this, CNT-based CA are promising for cell labeling and MRI tracking, being of great interest on stem cell-based therapies that have emerged as a promising approach for the treatment of different diseases. With this aim, Moghaddam et al. developed a method to coat the surface of gado-nanotubes (GNTs) with Gd and polyacrylic acid polymer generating a powerful MRI T1 contrast agent, with an extremely short T1 relaxation and an improved dispersibility in water without the need of surfactants [151]. The authors used these Reprinted from the journal

195

13



Topics in Current Chemistry (2020) 378:15

GNTs to safely label porcine bone marrow-derived mesenchymal stem cells that displayed excellent image contrast in phantom MRI. Our group reported an interesting and original application of SWCNTs to induce anisotropy in the diffusion of water molecules in a phantom [103]. This work offers new perspectives for contrast generation in diffusion tensor magnetic resonance imaging, a powerful MRI technique to explore the microstructure of healthy and pathological brain. The near-infrared radiation (NIR) absorption property of CNTs [152] can also be used to image their biodistribution in vivo. Yudasaka et al. [153] covered SWCNTs with a biocompatible polymer, which accumulated in brown fat, providing an imaging tool to visualize the distribution of this tissue in a preclinical model. Kim et al. [154] developed SWCNTs coated with gold and conjugated with antibody specific to the lymphatic vessel endothelial hyaluronan receptor to image the lymphatic vessels in mice. They induced a temperature increase of CNTs by NIR absorption using a laser beam and detected the nanotubes in the lymphatic vessels using photoacoustic and photothermal imaging. The NIR imaging techniques can also be combined with NIR guide photothermal therapy [155]. In this work, Liang et al. functionalized SWCNTs with polyethylene glycol to administrate them in BALB/c mice carrying 4T1 murine breast tumors in the inner knee. CNTs were directly injected on the primary tumor and visualized using photothermal imaging and MRI, not only in the tumor but also in the nearest metastatic lymph. CNTs can be functionalized with different radioisotopes like Y-86 [156], C-14 [157], I-125 [158], Tc-99 m [159] or Cu-64 [160], making them promising CAs for nuclear medicine approaches. Al-Jamal and coworkers [161] used single-photon emission computed tomography (SPECT)/computed tomography (CT) imaging to study in vivo the internalization of three different MWCNTs radiolabeled with 111In (Fig. 9). Furthermore, the particular features of carbon nanotubes make them good candidates to act as multimodal contrast agents with different imaging techniques. 4.2 Therapeutic Applications Nanoparticles have revolutionized the field of drug delivery in the last two decades. Currently used therapeutic drugs suffer from low selectivity and low half-life, making necessary it to give high doses to achieve the expected response, and thus increasing undesirable side-effects. A proper and selective drug delivery system would overcome most of these issues, and carbon nanotubes have been widely studied for this end [15]. As previously indicated, the nature of nanotubes permits a versatile chemistry allowing the attachment of drugs in a covalent or non-covalent manner with an efficient drug-loading capacity. Besides, these nanostructures can be appropriately functionalized with different hydrophilic molecules to specifically recognize the receptors overexpressed in the target cells, according to the alterations related to the specific pathology to be treated [162]. So, the optimal functionalization of CNTs make them promising drug-delivery systems in numerous therapies owing to their achieved biocompatibility, suitable size, ability to penetrate the cells, and

13

196

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

Fig. 9  Nano-SPECT/CT fused images of the whole body of a mouse. Images were recorded at 0.5, 3.5, and 24 h after injection of the 111In-DTPA-MWCNTs structures shown on the left side (with permission from Al-Jamal et al. [161])

their exceptional cell transfection capabilities [163, 164]. Hopeful studies have been reported to treat some of the most devastating pathologies nowadays, like cancer and neurodegenerative diseases, with CNT-based therapies using them either as drug carrier and delivery systems or taking advantage of nanotube physicochemical properties. 4.2.1 Cancer Treatment Nowadays, the main challenge of cancer therapies relies on the preparation of smart carriers able to accurately target the tumoral cells and to perform a controlled release of the therapeutic drug in response to the tumor microenvironment properties. The classic approach of nanotubes for cancer treatment is based on the transport and delivery of chemotherapeutic drugs that suffers from systemic toxicity, narrow therapeutic window, drug resistance, and low cellular penetration. CNTs provide a unique opportunity to improve the drug delivery to the tumor, enhancing Reprinted from the journal

197

13



Topics in Current Chemistry (2020) 378:15

local accumulation in the desired site and reducing the toxic effect of chemotherapy [165]. Relevant achievements in the field of nanotechnology to improve cancer treatment include carbon nanotubes loaded with classical therapeutic drugs and other anticancer agents like siRNA, chemosensitizers, radiosensitizers, and antiangiogenic compounds [166]. Liu et  al. [167] developed hyaluronic acid (HA)-modified amino SWVCNTs to target breast cancer cells overexpressing CD44, improving doxorubicin (DOX) deliver. In  vitro studies in MDA-MB-231 breast cancer cells showed an increased intracellular DOX delivery, higher inhibition of proliferation and induction of apoptosis, and decreased cell migration. The tubular structure of nanotubes is also optimal to attach DNA or RNA molecules to modulate the expression of specific genes [168]. In this sense, Taghavi et  al. [169] functionalized SWCNTs with AS1411 aptamer as ligand to target tumor cells and with DOX and small interfering RNA (siRNA) molecules to perform chemotherapy and gene therapy, respectively. This nanoplatform increased cell death in a model of human gastric cancer measured in vitro when compared to the individual CNT-based treatments or with free DOX. Chang Guo et al. [170] prepared cationic MWCNT-NH3+ to deliver siRNA against polo-like kinase 1 (PLK1) in a lung carcinoma model in  vivo. CNTs were injected directly in the tumoral mass, finding higher efficiency of PLK1 silencing compared to liposomes due to a higher CNT cell penetration. The possibilities that carbon nanotubes offers as drug carriers in the treatment of malignant tumors is very large [171–173], but they afford other potential approaches in cancer treatment, like photothermal therapy (PTT) [174]. CNTs can be used as external agents in PTT because they absorb NIR radiation and efficiently convert it into heat energy, allowing the ablation of the cells in the CNT surroundings. Virani et  al. used this phenomenon to target bladder cancer cells [175]. They conjugated SWCNTs with annexin V, which specifically binds to bladder cancer cells and tested the method in vivo. Once the CNTs reached the cancer, the tumor was heated using NIR light, inducing cell death and preserving the healthy bladder wall. They founded no damage on the bladder 24 h after treatment and no tumors were visible. To date, the main approach for PTT with CNTs has been based on the irradiation by laser or infrared light, limiting the application to superficial tissues due to low penetration capacity, but CNTs can also be exposed to external electromagnetic fields in a non-contact manner, achieving higher penetration while obtaining the desired thermal effects [176]. Nowadays, chemoresistant tumors with a high tumor load and in advanced stages require the combination of different therapies for effective treatment. CNTs are well suited for this, with carbon-based nanoparticles being used for combining thermal therapy with other therapeutic approaches based on drug delivery [177]. Wang et  al. reported the use of NIR photothermal therapy and RNAi to enhanced tumor cells death in a prostate cancer model [178]. They synthesized SWCNTs functionalized with polyethylenimine to allow siRNA complexation, and decorated the CNT with a peptide motif that specifically binds to tumor cells. The cancer therapy effect of these CNTs was tested in  vitro and in  vivo using PC-3 tumor cells. The CNTs silenced the target gen, causing significant inhibition

13

198

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



of the tumor growth, and, when combined with photothermal therapy, the results showed a more effective treatment. 4.2.2 Neurodegeneration Treatment One of the big problems in the treatment of neurodegenerative diseases like Parkinson, dementia, or Alzheimer’s, is the presence of the blood–brain barrier (BBB). The BBB restricts the passage of bacteria and toxic molecules to avoid brain damage but also prevents the brain uptake of most pharmaceuticals, precluding the pass of drugs from blood circulation to brain tissue. In fact, neurological disorders belong to the group of pathologies with the greatest need of new technologies to improve diagnosis and provide an adequate therapy, and nanobiotechnology can help in both. Al-Jamal’s laboratory proved [179–181] that CNTs are able to cross the blood–brain barrier without damaging it. They used in  vitro cultures of primary porcine brain endothelial cells, which simulate the BBB, showing that MWCNTs functionalized with N ­ H3 crossed 13% of the layer in 72 h. They also demonstrated in vivo that 5 min after intravenous injection of MWCNTs-NH3 in mice, 1% of the injected dose crossed the BBB. In further experiments, they studied the impact of the diameter of the nanotube and the functionalization with angiopep-2 (AGP), a ligand for the low-density lipoprotein receptor related protein-1, in the extent of passing the BBB. They found that wider nanotubes linked with AGP showed a higher effect on crossing the BBB than those not functionalized. Thinner CNTs did not depict any improvement in the brain uptake between functionalized and unfunctionalized MWCNTs, although they exhibited more uptake than the wider ones. Finally, they tested them in vivo using a GL261 glioma model, showing higher uptake for CNT functionalized with AGP. Costa et al. [182] took advantage of this property to develop an Alzheimer’s biomarker based on gadolinium. MWCNTs were functionalized with the Pittsburgh Compound B, a molecule that binds to the amyloid beta plaques present in Alzheimer’s disease and is currently used for imaging Alzheimer’s with PET. The group studied the in vivo distribution of the nanotubes, finding that they were able to cross the BBB, making them a good contrast agent to detect the presence of the plaques with MRI. Lohan et al. [183] reported the development of MWCNTs functionalized with berberine (BRB), an isoquinoline alkaloid used in dementia and other mental disorders, to evaluate the anti-Alzheimer’s potential of the drug. The authors injected the nanotubes in healthy male Wistar rats after intracerebroventricular administration of amyloid beta peptide. They proved that BRBMWCNTS crossed the BBB, reaching the plaques, and found a decrease of malondialdehyde and reduced glutathione, two of Alzheimer’s biomarkers. The structural damage occurring in some neuro-diseases can also be minimized with carbon nanotubes. Hassanzadeh et  al. [184] prepared MWCNTs-NGF complexed with nerve growth factor (NGF), a protein related to the survival and maintenance of neurons population, and investigated its effect in an in  vitro model of ischemia. The results were compared with those obtained with free NGF, finding higher protection for the new CNT-NGF complex. As in cancer, CNTs found further applications in neurodegenerative diseases beyond drug delivery. CNTs can be used to activate brain cells due to their electrical Reprinted from the journal

199

13

Topics in Current Chemistry (2020) 378:15

properties and tubular shape. Silviana Fiorito et al. [185] reported the activation of microglia using MWCNTs with highly electro-conductive properties, polarizing the microglia cells into M1 state (pro-inflammatory) after 24 h of cell exposure and M2 state (anti-inflammatory) after 48 h of exposition. They detected this phenomenon in vitro using primary rat microglial cells, suggesting that the electro-stimulation of these CNTs could modulate microglial responses. The use of carbon nanotubes as scaffolds for neurons has also been well studied, as they are good platforms for protecting neuron damage in ischemia [186]. Jung Lee et al. pretreated rats with amine-modified SWCNTs before an ischemic surgery, finding less tissue damage and better motor function in the treated animals after the insult. More recently, CNTs were reported as good scaffolds for different types of neurons, like retinal neurons or sciatic nerve [187]. In this work, the authors studied retinal cell growth in  vitro using MWCNTs as frame to these cells. They followed the procedure with rat and human retinal cells, finding an efficient growth for both of them. Salehi et al. [188] developed a conduit produced from polylactic acid, MWCNTs, and gelatin nanofibrils coated with the recombinant human erythropoietin-loaded chitosan nanoparticles. They loaded this conduit with Schwann cells and implanted it in rats with sciatic nerve defect, finding improved recovery rate. Finally, carbon nanotubes have also shown some neuroprotector properties. Xue et al. [189] reported that C57BL/6J mice pretreated with aggregated SWCNTs significantly inhibited self-administration of methamphetamine. The electrochemical assays indicated that nanotubes made the oxidation of extracellular dopamine in the striatum easier, suggesting the potential use of aggregated SWCNTs for the treatment of methamphetamine addiction. 4.3 Theranostic Applications Theranostic compounds are nanoparticles integrating both diagnostic and therapeutic capabilities into a platform. CNTs have been intensively studied for theranostic applications [163], as they can be easily coupled with different molecules/structures achieving synergetic analytic and curative effects [190]. Basically, this approach relies on the CNT functionalization with structures carrying abilities for detecting and treating the disease, achieving at the same time cell targeting and delivering of the drug under microenvironmental stimuli. Some examples have already been presented in earlier. Mashal et al. [191] constructed tissue-mimicking materials and found that SWCNTs could enhance dielectric contrast between tumoral and healthy tissue for microwave detection and hyperthermia treatment in breast cancer. More recently, Al Faraj et al. [192] described the preparation and the in vitro and in vivo characterization of drug-conjugated SWCNTs as nanocarriers for breast cancer therapy. The authors aimed to target a stem cell subpopulation with a combination of therapeutic drugs (paclitaxel and salinomycin) selectively delivered via biocompatible CD44 antibody-conjugated SWCNTs with a pH-responsive release. The therapy was non-invasively monitored in tumor-bearing mice with MRI and bioluminescence imaging (BLI), and the results clearly showed enhanced therapeutic effect of

13

200

Reprinted from the journal

Fig. 10  In vivo drug-conjugated SWCNT therapeutic efficacy assessed using BLI and MRI: a schematic diagram illustrating the functionalized SWCNTs as theranostic carriers, b representative bioluminescence images, c representative axial MR images, and d quantitative measurements of tumor volume in MDA-MB-231 tumor-bearing mice. Imaging protocols were performed pre- (t = 0) and up to 35 days post-treatment (modified with permission from Al Faraj et al. [192])

Topics in Current Chemistry (2020) 378:15

Reprinted from the journal



201

13



Topics in Current Chemistry (2020) 378:15

the combined therapy compared to treatment with individual drug-conjugated nanocarriers or free drug suspensions (Fig. 10). CNTs can also be coupled to other nanosystems with this purpose, such as magnetic particles, quantum dots, nanocomposites, etc. Zhang et al. [193] functionalized magnetofluorescent MWCNTs for MRI/fluorescence imaging by adding a Gd complex doped with quantum dots. They also combined the nanotubes with doxorubicin to be used as a chemotherapeutic agent. The nanotubes were intratumorally injected in a preclinical model of adenocarcinoma, and visualized with photoacoustic imaging and magnetic resonance T1-weighted images. Using NIR absorption properties, the authors carried out photothermal therapy in combination with selective release of DOX in the tumor. In a similar way, Hou et  al. [194] functionalized SWCNTs with HA, DOX, and gadolinium to visualize with MRI the biodistribution of the nanotubes and their antitumor effect in a preclinical model of breast cancer. CNTs can also be used as theranostic dual contrast agents for fluorescence/MRI and photothermal therapy, all in a system [195]. The authors So, Zhang, and colleagues reported the preparation of and in  vitro and in  vivo characterization of MWCNT-magnetofluorescent carbon quantum dots/DOX nanocomposites. They validated the diagnostic and therapeutic capacities of the compounds with MRI and fluorescent methodologies in adenocarcinoma cells and a pulmonary cancer mouse model. The study confirmed that the platform is able to target cancer cells and deliver drugs intracellularly upon NIR irradiation, achieving the effective elimination of the tumors through chemo/photothermal synergistic therapeutic effect. In a similar way, Xiaojing Wang et al. [196] described the synthesis of modified DNASWCNTs with gold-decorated nanoparticles and a surface modification with PEG, which achieve the selective photothermal ablation of cancer cells. These SWNT-AuPEG nanocomposites are optical theranostic probes for cancer treatment by PPT and detectable by Raman spectroscopic imaging. Another study, carried out by Zhao and coworkers [197], reports the coating of CNTs with polydopamine and further modification by PEG to chelate manganese, achieving contrast in MRI, and to label 131I enabling nuclear imaging and radioisotope cancer therapy. The system was tested in  vitro and in a mouse model of breast cancer, assessing radioisotope therapy in combination with NIR-triggered photothermal treatment. Results revealed efficient tumor accumulation of the nanotubes and confirmed a remarkable synergistic antitumor therapeutic effect. 4.4 Tissue Engineering Applications The main clue in regeneration and construction of tissues is the development of a suitable biological scaffold. In this context, owing to their unique characteristics and properties, carbon nanotubes are emerging as smart nanomaterials for tissue engineering purposes. They are valued as ideal structures that can support and boost the growth and proliferation of many different tissues [15].

13

202

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



4.4.1 CNTs and Cell Growing Nanobiotechnology can have a great impact in the management of nervous system pathologies by developing optimal structures for neural prosthetic applications. The design of biocompatible implants for neuron repair/regeneration ideally requires high cell adhesion as well as good electrical conductivity. Carbon nanotubes entail all of these requirements, including high binding affinity and excellent electrical conductivity, making them ideal materials for neuro-implant development aimed to grow neurons and repair neuronal damage. Some examples have already presented in earlier. Lovan et al. [198] showed that carbon nanotubes possess a good surface for supporting dendrite elongation and cell adhesion. Experiments were carried out on neonatal hippocampal neuron networks cultured on dispersed MWCNTs. Results suggested that the growth of neuronal circuits on a nanotube grid is accompanied by a significant increase in network activity, probably due to the high electrical conductivity of these nanomaterials. In the same line, Mazzatenta et al. [199] described the preparation of an integrated SWCNT–neuron system by growing hippocampal cells on the CNTs. Theoretical and experimental results indicated that SWCNTs can stimulate the activity of the brain circuits. However, CNT applications in tissue regeneration or engineering expand far beyond nervous tissue. Ren et al. [200] used CNTs to develop artificial myocardial tissue where cell growth was aided by the conductivity of oriented CNTs. Special nanotube geometry also beneficiated artificial bone generation, making CNTs a very suitable bone scaffold both in vitro and in vivo [201, 202]. Finally, CNTs make an important contribution in the generation of synthetic fiber muscle [203, 204]. 4.4.2 CNT‑Based Hydrogels Although conventional hydrogels are biocompatible and suitable for culturing or fabricating different cell types and tissues, their low mechanical strength and lack of electrical conductivity have limited their biomedical applications for skeletal muscles, cardiac and neural cells. Nevertheless, the development of hybrid nanocomposite systems can overcome these limitations enabling the preparation of bioscaffolds with tunable electrical and mechanical features. In fact, in the last few years, CNTbased hybrid hydrogels are emerging as innovative candidates with applications in regenerative medicine and tissue engineering [205]. Shin et al. [206] prepared different nanotube–hydrogel hybrid systems that showed significantly improved electrophysiological and mechanical properties. The authors seeded neonatal rat cardiomyocytes onto these hybrid MWCNTs hydrogels, obtaining functional cardiac patches that showed excellent mechanical integrity and advanced electrophysiological functions. Another study reports the addition of functionalized MWCNTs to alginate to generate composite hydrogels improving the mechanical, physical, and biological features compared with the starting materials [207]. The obtained hybrid MWCNT–alginate gels were porous, showed less degradation, enhanced HeLa cells adhesion and had greater cell proliferation, proving the potential utility of these structures as novel substrates for tissue preparation. Sun et  al. [208] incorporated SWCNTs into collagen hydrogels, which improved cell alignment and assembly, Reprinted from the journal

203

13



Topics in Current Chemistry (2020) 378:15

leading to the formation of engineered cardiac tissues with stronger contraction capacity. The in vitro studies suggested that hybrid SWCNT/collagen hydrogels can be promising tissue scaffolds for cardiac regeneration after myocardial arrest. In a similar line, a recent study described a biohybrid hydrogel prepared from hydrazidefunctionalized CNTs and solubilized pericardial matrix. This hydrogel is a suitable environment for maturation of human-induced pluripotent stem cell-derived cardiomyocytes, constituting a promising material for stem cell-based cardiac tissue engineering [209]. 4.4.3 CNTs and Stem Cells Stem cells (SC) have attracted the attention of researchers around the world during the last decade due to their ability to self-renew and differentiate, depicting multiple potential applications for tissue engineering and regenerative medicine. CNTs presented excellent properties as culture substrate [210], having the ability to dynamically direct the SC lineage, modulating proliferation and differentiation of various types of SC. Different studies have reported the improvements achieved by growing SC in CNT-based nanocomposites. These materials were able to promote differentiation of mouse neural SC to neurons and oligodendrocytes [211], neural differentiation of human embryonic stem cells [212], and differentiation of human mesenchymal stem cells [213], among others. 4.5 Other Applications The versatility of carbon nanotubes in biomedicine and preclinical setup is very large, with many studies reporting the use of CNTs to solve the problem under investigation. Carbon nanotubes are being used, for example, to construct brain electrodes [214], or as biosensors for bone cells [215] or the influenza virus [112]. By linking CNTs to antigenic peptides, they could improve the major hurdles associated with vaccine delivery [216]. In fact, they can act as excellent vaccine carrier systems having a great potential to stimulate the innate immune response. Also, the combination of CNTs with antibiotics could help to overcome the growing problem of antibiotics resistance [190]. When properly functionalized, CNTs can also be used in antiviral drug delivery [217].

5 Conclusions Bionanotechnology is a new promising tool for improving the management of numerous pathologies. A wide range of nanomaterials have been prepared, characterized, and evaluated for numerous biomedical applications to solve specific problems not achievable with classical approaches. Among them, carbon nanotubes present unique features and physical, chemical, and biological properties. These intrinsic attributes can be notably improved though coating, surface functionalization, and decoration with different molecules, ligands, or nanostructures. All of

13

204

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



these turn CNTs into ideal platforms with diagnostic, therapeutic, and theranostic possibilities not accessible for other nanoparticles and with promising preclinical and clinical applications. In this review, the biomedical applications of CNTs have been addressed in different fields of drug and/or gene delivery, bioimaging, and tissue engineering. The special electronic and mechanical qualities of nanotubes make them suitable for theranostic applications as providing an exceptional platform for combining disease detection and treatment capacity at the same time. These structures can be used in diagnosis as efficient biosensors or contrast agent for non-invasive imaging, and at the same time, increase the lifetime of drugs in organism and facilitate their direct delivery within cells of a target-specific tissue. Also, nanotubes have reached a vital relevance in biotechnology, being an excellent scaffold on their own or taking the part of hybrid materials for regenerative medicine. They have proved to be useful in cell growth and proliferation, enabling the engineering of different tissues. The studies included in this review have highlighted the potential of CNTs and justified all the efforts in optimizing their use as an alternative therapy for complicated medical conditions with no current treatments. Nevertheless, despite the numerous and intensive studies, some important inconveniences, which preclude potential clinical applications, need to be overcome, with toxicity probably being the most important handicap. The triad of CNTs functionalization, type, size, and purity have been identified as leading causes determining the utility of nanotube-based complex in vivo. Collating reliable cellular and animal data with respect to molecularly well-defined architectures provides a basis for further breakthroughs on the horizon. Acknowledgements  This study was funded by grants from the Ministry of Economy, Industry and Competitivity (SAF2017-83043-R), and by the Program MULTITARGET&VIEW-CM from Community of Madrid, Spain (S2017/BMD-3688), involving contributions from FEDER and FSE funds. Authors Contributions  Pilar López-Larrubia had the idea for the article. Viviana Negri, Jesús PachecoTorres, Daniel Calle, and Pilar Lopez-Larrubia performed the literature search and data analysis, drafted and critically revised the work.

Compliance with Ethical Standards  Conflict of interest  The authors declare that they have no conflicts of interest.

References 1. Ahsan MA, Jabbari V, Islam MT, Turley RS, Dominguez N, Kim H, Castro E, Hernandez-Viezcas JA, Curry ML, Lopez J, Gardea-Torresdey JL, Noveron JC (2019) Sustainable synthesis and remarkable adsorption capacity of MOF/graphene oxide and MOF/CNT based hybrid nanocomposites for the removal of Bisphenol A from water. Sci Total Environ 673:306–317. https​://doi. org/10.1016/j.scito​tenv.2019.03.219 2. Ahsan MA, Jabbari V, Imam MA, Castro E, Kim H, Curry ML, Valles-Rosales DJ, Noveron JC (2020) Nanoscale nickel metal organic framework decorated over graphene oxide and carbon

Reprinted from the journal

205

13



Topics in Current Chemistry (2020) 378:15

nanotubes for water remediation. Sci Total Environ 698:134214. https​://doi.org/10.1016/j.scito​ tenv.2019.13421​4 3. Cova CM, Zuliani A, Santiago ARP, Caballero A, Muñoz-Batista MJ, Luque R (2018) Microwaveassisted preparation of Ag/Ag2S carbon hybrid structures from pig bristles as efficient HER catalysts. J Mater Chem A 6:21516–21523 4. Franco A, Cebrián-García S, Rodríguez-Padrón D, Puente-Santiago AR, Muñoz-Batista MJ, Caballero A, Balu AM, Romero AA, Luque R (2018) Encapsulated laccases as effective electrocatalysts for oxygen reduction reactions. ACS Sustain Chem Eng 6:11058–11062 5. Ahsan MA, Jabbari V, El-Gendy AA, Curry ML, Noveron JC (2019) Ultrafast catalytic reduction of environmental pollutants in water via MOF-derived magnetic Ni and Cu nanoparticles encapsulated in porous carbon. Appl Surf Sci 597:142608 6. Ahsan MA, Deemer E, Fernandez-Delgado O, Wang H, Curry ML, El-Gendy AA, Noveron JC (2019) Fe nanoparticles encapsulated in MOF-derived carbon for the reduction of 4-nitrophenol and methyl orange in water. Catal Commun 130:105753 7. Ahsan MA, Fernandez-Delgado O, Deemer E, Wang H, El-Gendy AA, Curry ML, Noveron JC (2019) Carbonization of Co-BDC MOF results in magnetic C@Co nanoparticles that catalyze the reduction of methyl orange and 4-nitrophenol in water. J Mol Liq 290:111059 8. Ostovar S, Franco A, Puente-Santiago AR, Pinilla-de Dios M, Rodriguez-Padron D, Shaterian HR, Luque R (2018) Efficient mechanochemical bifunctional nanocatalysts for the conversion of isoeugenol to vanillin. Front Chem 6:77. https​://doi.org/10.3389/fchem​.2018.00077​ 9. Rodríguez-Padrón D, Puente-Santiago AR, Balu AM, Muñoz-Batista MJ, Luque R (2019) Environmental catalysis: present and future. ChemCatChem 11:18–38 10. Pacheco-Torgal F, Diamanti MV, Nazari A, Goran-Granqvist C, Pruna A, Amirkhanian SE (2018) Nanotechnology in eco-efficient construction: materials, processes and applications. Woodhead Publishing, Sawston 11. Sanad MF, Shalan AE, Bazid SM, Serea ESA, Hashem EM, Nabih S, Ahsan MA (2019) A graphene gold nanocomposite-based 5-FU drug and the enhancement of the MCF-7 cell line treatment. RSC Adv 9:31021–31029 12. Chen S, Li R, Li X, Xie J (2018) Electrospinning: an enabling nanotechnology platform for drug delivery and regenerative medicine. Adv Drug Deliv Rev 132:188–213. https​://doi.org/10.1016/j. addr.2018.05.001 13. Yang Y, Chawla A, Zhang J, Esa A, Jang HL, Khademhosseini A (2019) Applications of nanotechnology for regenerative medicine; healing tissues at the nanoscale. Principles of Regenerative Medicine, pp 485–504 14. Iijima S (1991) Helical microtubules of graphitic carbon. Nature 354(6348):56–58 15. Simon J, Flahaut E, Golzio M (2019) Overview of carbon nanotubes for biomedical applications. Materials (Basel). https​://doi.org/10.3390/ma120​40624​ 16. Eatemadi A, Daraee H, Karimkhanloo H, Kouhi M, Zarghami N, Akbarzadeh A, Abasi M, Hanifehpour Y, Joo SW (2014) Carbon nanotubes: properties, synthesis, purification, and medical applications. Nanoscale Res Lett 9(1):393. https​://doi.org/10.1186/1556-276x-9-393 17. Galano A (2010) Carbon nanotubes: promising agents against free radicals. Nanoscale 2(3):373–380 18. Lu JP (1997) Elastic properties of carbon nanotubes and nanoropes. Phys Rev Lett 79(7):1297–1300 19. Treacy MMJ, Ebbesen TW, Gibson JM (1996) Exceptionally high Young’s modulus observed for individual carbon nanotubes. Nature 381(6584):678–680 20. Wong EW, Sheehan PE, Lieber CM (1997) Nanobeam mechanics: elasticity, strength, and toughness of nanorods and nanotubes. Science 277(5334):1971–1975 21. Wei BQ, Vajtai R, Ajayan PM (2001) Reliability and current carrying capacity of carbon nanotubes. Appl Phys Lett 79(8):1172–1174 22. White CT, Todorov TN (1998) Carbon nanotubes as long ballistic conductors. Nature 393(6682):240–242 23. Ebbesen TW, Lezec HJ, Hiura H, Bennett JW, Ghaemi HF, Thio T (1996) Electrical conductivity of individual carbon nanotubes. Nature 382(6586):54–56 24. Hills G, Lau C, Wright A, Fuller S, Bishop MD, Srimani T, Kanhaiya P, Ho R, Amer A, Stein Y, Murphy D, Arvind Chandrakasan A, Shulaker MM (2019) Modern microprocessor built from complementary carbon nanotube transistors. Nature 572(7771):595–602. https​://doi.org/10.1038/ s4158​6-019-1493-8

13

206

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



25. Wan X, Dong J, Xing DY (1998) Optical properties of carbon nanotubes. Phys Rev B 58:4. https​:// doi.org/10.1103/PhysR​evB.58.6756 26. Hone JL, Biercuk MJ, Johnson AT, Batlogg B, Benes Z, Fisher JE (2002) Thermal properties of carbon nanotubes and nanotube-based materials. Appl Phys A 74:5. https​://doi.org/10.1007/s0033​ 90201​277 27. Ardavan A, Austwick M, Benjamin SC, Briggs GAD, Dennis TJS, Ferguson A, Hasko DG, Kanai M, Khlobystov AN, Lovett BW, Morley GW, Oliver RA, Pettifor DG, Porfyrakis K, Reina JH, Rice JH, Smith JD, Taylor RA, Williams DA, Adelmann C, Mariette H, Hamers RJ (2003) Nanoscale solid-state quantum computing. Philos Trans R Soc A 361(1808):1473–1485. https​:// doi.org/10.1098/rsta.2003.1214 28. Monthioux M, Flahaut E (2007) Meta- and hybrid-CNTs: a clue for the future development of carbon nanotubes. Mater Sci Eng C 27(5):1096–1101. https​://doi.org/10.1016/j.msec.2006.07.032 29. Chae S, Kim D, Lee K-j, Lee D, Kim Y-O, Jung YC, Rhee SD, Kim KR, Lee J-O, Ahn S, Koh B (2018) Encapsulation and enhanced delivery of topoisomerase I inhibitors in functionalized carbon nanotubes. ACS Omega 3(6):5938–5945. https​://doi.org/10.1021/acsom​ega.8b003​99 30. Tasis D, Tagmatarchis N, Bianco A, Prato M (2006) Chemistry of carbon nanotubes. Chem Rev 106(3):1105–1136 31. Joseph S, Mashl RJ, Jakobsson E, Aluru NR (2003) Electrolytic transport in modified carbon nanotubes. Nano Lett 3(10):1399–1403. https​://doi.org/10.1021/nl034​6326 32. Pan X, Bao X (2011) The effects of confinement inside carbon nanotubes on catalysis. Acc Chem Res 44(8):553–562. https​://doi.org/10.1021/ar100​160t 33. Mirabootalebi SO, Akbari G (2017) Methods for synthesis of carbon nanotubes—review 34. Kim SWKT, Kim YS, Choi HS, Lim HJ, Yang SJ, Park CR (2012) Surface modifications for the effective dispersion of carbon nanotubes in solvents and polymers. Carbon 50:3–33 35. Wu HC, Chang X, Liu L, Zhao Y (2010) Chemistry of carbon nanotubes in biomedical applications. J Mater Chem A 20:1036–1052 36. Britz DA, Khlobystov AN (2006) Noncovalent interactions of molecules with single-walled carbon nanotubes. Chem Soc Rev 35(7):637–659. https​://doi.org/10.1039/b5074​51g 37. Ajayan PM, Iijima S (1993) Capillarity-induced filling of carbon nanotubes. Nature 361(6410):333–334. https​://doi.org/10.1038/36133​3a0 38. Dujardin E, Ebbesen TW, Hiura H, Tanigaki K (1994) Capillarity and wetting of carbon nanotubes. Science 265(5180):1850–1852. https​://doi.org/10.1126/scien​ce.265.5180.1850 39. Campo J, Piao Y, Lam S, Stafford CM, Streit JK, Simpson JR, Hight Walker AR, Fagan JA (2016) Enhancing single-wall carbon nanotube properties through controlled endohedral filling. Nanoscale Horizons 1(4):317–324. https​://doi.org/10.1039/C6NH0​0062B​ 40. Li J, Yap SQ, Chin CF, Tian Q, Yoong SL, Pastorin G, Ang WH (2012) Platinum(iv) prodrugs entrapped within multiwalled carbon nanotubes: selective release by chemical reduction and hydrophobicity reversal. Chem Sci 3(6):2083–2087. https​://doi.org/10.1039/C2SC0​1086K​ 41. Karousis N, Tagmatarchis N, Tasis D (2010) Current progress on the chemical modification of carbon nanotubes. Chem Rev 110(9):5366–5397. https​://doi.org/10.1021/cr100​018g 42. Jeon IY, Chang DW, Kumar NA, Baek JB (2011) Functionalization of carbon nanotubes. Carbon Nanotubes Polymer Nanocomposites. https​://doi.org/10.5772/979 43. Jha N, Ramesh P, Bekyarova E, Tian XJ, Wang FH, Itkis ME, Haddon RC (2013) Functionalized single-walled carbon nanotube-based fuel cell benchmarked against US DOE 2017 Technical Targets. Sci Rep-Uk. https​://doi.org/10.1038/srep0​2257 44. Dyke CA, Tour JM (2004) Covalent functionalization of single-walled carbon nanotubes for materials applications. J Phys Chem A 108(51):11151–11159. https​://doi.org/10.1021/jp046​274g 45. Hirsch A (2002) Functionalization of single-walled carbon nanotubes. Angew Chem Int Ed 41(11):1853–1859 46. Singh P, Campidelli S, Giordani S, Bonifazi D, Bianco A, Prato M (2009) Organic functionalisation and characterisation of single-walled carbon nanotubes. Chem Soc Rev 38(8):2214–2230. https​://doi.org/10.1039/b5181​11a 47. Alshehri R, Ilyas AM, Hasan A, Arnaout A, Ahmed F, Memic A (2016) Carbon nanotubes in biomedical applications: factors, mechanisms, and remedies of toxicity. J Med Chem 59(18):8149– 8167. https​://doi.org/10.1021/acs.jmedc​hem.5b017​70 48. Hamwi A, Alvergnat H, Bonnamy S, Beguin F (1997) Fluorination of carbon nanotubes. Carbon 35(6):723–728

Reprinted from the journal

207

13



Topics in Current Chemistry (2020) 378:15

49. Mickelson ET, Huffman CB, Rinzler AG, Smalley RE, Hauge RH, Margrave JL (1998) Fluorination of single-wall carbon nanotubes. Chem Phys Lett 296(1–2):188–194. https​://doi.org/10.1016/ s0009​-2614(98)01026​-4 50. Kelly KF, Chiang IW, Mickelson ET, Hauge RH, Margrave JL, Wang X, Scuseria GE, Radloff C, Halas NJ (1999) Insight into the mechanism of sidewall functionalization of single-walled nanotubes: an STM study. Chem Phys Lett 313(3–4):445–450. https​://doi.org/10.1016/S0009​ -2614(99)00973​-2 51. Touhara H, Inahara J, Mizuno T, Yokoyama Y, Okanao S, Yanagiuch K, Mukopadhyay I, Kawasaki S, Okino F, Shirai H, Xu WH, Kyotani T, Tomita A (2002) Property control of new forms of carbon materials by fluorination. J Fluorine Chem 114(2):181–188. https​://doi.org/10.1016/s0022​ -1139(02)00026​-x 52. Stevens JL, Huang AY, Peng H, Chiang IW, Khabashesku VN, Margrave JL (2003) Sidewall amino-functionalization of single-walled carbon nanotubes through fluorination and subsequent reactions with terminal diamines. Nano Lett 3(3):331–336. https​://doi.org/10.1021/nl025​944w 53. Zhang L, Kiny VU, Peng H, Zhu J, Lobo RFM, Margrave JL, Khabashesku VN (2004) Sidewall functionalization of single-walled carbon nanotubes with hydroxyl group-terminated moieties. Chem Mater 16(11):2055–2061. https​://doi.org/10.1021/cm035​349a 54. Boul PJ, Liu J, Mickelson ET, Huffman CB, Ericson LM, Chiang IW, Smith KA, Colbert DT, Hauge RH, Margrave JL, Smalley RE (1999) Reversible sidewall functionalization of buckytubes. Chem Phys Lett 310(3–4):367–372 55. Saini RK, Chiang IW, Peng HQ, Smalley RE, Billups WE, Hauge RH, Margrave JL (2003) Covalent sidewall functionalization of single-wall carbon nanotubes. J Am Chem Soc 125(12):3617–3621 56. Unger E, Graham A, Kreupl F, Liebau M, Hoenlein W (2002) Electrochemical functionalization of multi-walled carbon nanotubes for solvation and purification. Curr Appl Phys 2(2):107–111 57. Amirian M, Nabipour Chakoli A, Sui JH, Cai W (2012) Enhanced mechanical and photoluminescence effect of poly(l-lactide) reinforced with functionalized multiwalled carbon nanotubes. Polym Bull 68(6):1747–1763. https​://doi.org/10.1007/s0028​9-012-0700-7 58. Chakoli AN, He JM, Huang YD (2018) Collagen/aminated MWCNTs nanocomposites for biomedical applications. Mater Today Commun 15:128–133. https​://doi.org/10.1016/j.mtcom​ m.2018.03.003 59. Graupner R, Abraham J, Wunderlich D, Vencelová A, Lauffer P, Röhrl J, Hundhausen M, Ley L, Hirsch A (2006) Nucleophilic–alkylation–reoxidation: a functionalization sequence for single-wall carbon nanotubes. J Am Chem Soc 128(20):6683–6689. https​://doi.org/10.1021/ja060​7281 60. Dyke CA, Stewart MP, Maya F, Tour JM (2004) Diazonium-based functionalization of carbon nanotubes: XPS and GC–MS analysis and mechanistic implications. Synlett 1:155–160. https​://doi. org/10.1055/s-2003-44983​ 61. Bahr JL, Yang J, Kosynkin DV, Bronikowski MJ, Smalley RE, Tour JM (2001) Functionalization of carbon nanotubes by electrochemical reduction of aryl diazonium salts: a bucky paper electrode. J Am Chem Soc 123(27):6536–6542. https​://doi.org/10.1021/ja010​462s 62. Strano MS, Dyke CA, Usrey ML, Barone PW, Allen MJ, Shan H, Kittrell C, Hauge RH, Tour JM, Smalley RE (2003) Electronic structure control of single-walled carbon nanotube functionalization. Science 301(5639):1519–1522. https​://doi.org/10.1126/scien​ce.10876​91 63. Dyke CA, Tour JM (2003) Unbundled and highly functionalized carbon nanotubes from aqueous reactions. Nano Lett 3(9):1215–1218. https​://doi.org/10.1021/nl034​537x 64. Leventis HC, Wildgoose GG, Davies IG, Jiang L, Jones TG, Compton RG (2005) Multiwalled carbon nanotubes covalently modified with fast black K. ChemPhysChem 6(4):590–595. https​://doi. org/10.1002/cphc.20040​0536 65. Bahr JL, Tour JM (2001) Highly functionalized carbon nanotubes using in  situ generated diazonium compounds. Chem Mater 13(11):3823–3824. https​://doi.org/10.1021/cm010​9903 66. Mitchell CA, Bahr JL, Arepalli S, Tour JM, Krishnamoorti R (2002) Dispersion of functionalized carbon nanotubes in polystyrene. Macromolecules 35(23):8825–8830. https​://doi.org/10.1021/ ma020​890y 67. Hudson JL, Casavant MJ, Tour JM (2004) Water-soluble, exfoliated, nonroping single-wall carbon nanotubes. J Am Chem Soc 126(36):11158–11159. https​://doi.org/10.1021/ja046​7061 68. Dyke CA, Tour JM (2003) Solvent-free functionalization of carbon nanotubes. J Am Chem Soc 125(5):1156–1157. https​://doi.org/10.1021/ja028​9806

13

208

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15 69. Kooi SE, Schlecht U, Burghard M, Kern K (2002) Electrochemical modification of single carbon nanotubes. Angew Chem Int Ed Engl 41(8):1353–1355 70. Balasubramanian K, Sordan R, Burghard M, Kern K (2004) A selective electrochemical approach to carbon nanotube field-effect transistors. Nano Lett 4(5):827–830. https​://doi.org/10.1021/nl049​ 806d 71. Delgado JL, de la Cruz P, Langa F, Urbina A, Casado J, Navarrete JTL (2004) Microwave-assisted sidewall functionalization of single-wall carbon nanotubes by Diels–Alder cycloaddition. Chem Commun 15:1734–1735 72. Araújo RF, Proença MF, Silva CJ, Castro TG, Melle-Franco M, Paiva MC, Villar-Rodil S, Tascón JMD (2016) Grafting of adipic anhydride to carbon nanotubes through a Diels–Alder cycloaddition/oxidation cascade reaction. Carbon 98:421–431. https​://doi.org/10.1016/j.carbo​n.2015.11.004 73. Georgakilas V, Kordatos K, Prato M, Guldi DM, Holzinger M, Hirsch A (2002) Organic functionalization of carbon nanotubes. J Am Chem Soc 124(5):760–761. https​://doi.org/10.1021/ja016​ 954m 74. Georgakilas V, Tagmatarchis N, Pantarotto D, Bianco A, Briand JP, Prato M (2002) Amino acid functionalisation of water soluble carbon nanotubes. Chem Commun 24:3050–3051. https​://doi. org/10.1039/b2098​43a 75. Lu X, Tian F, Xu X, Wang NQ, Zhang Q (2003) Theoretical exploration of the 1,3-dipolar cycloadditions onto the sidewalls of (n, n) armchair single-wall carbon nanotubes. J Am Chem Soc 125(34):10459–10464. https​://doi.org/10.1021/ja034​662a 76. Yao ZL, Braidy N, Botton GA, Adronov A (2003) Polymerization from the surface of singlewalled carbon nanotubes—preparation and characterization of nanocomposites. J Am Chem Soc 125(51):16015–16024. https​://doi.org/10.1021/ja037​564y 77. Calcio Gaudino E, Tagliapietra S, Martina K, Barge A, Lolli M, Terreno E, Lembo D, Cravotto G (2014) A novel SWCNT platform bearing DOTA and β-cyclodextrin units. “One shot” multidecoration under microwave irradiation. Org Biomol Chem 12(26):4708–4715. https​://doi.org/10.1039/ c4ob0​0611a​ 78. Tobias G, Mendoza E, Ballesteros B (2012) Functionalization of Carbon Nanotubes. In: Bhushan B (ed) Encyclopedia of nanotechnology. Springer Netherlands, Dordrecht, pp 911–919. https​://doi. org/10.1007/978-90-481-9751-4_48 79. Zhang ZY, Xu XC (2015) Nondestructive covalent functionalization of carbon nanotubes by selective oxidation of the original defects with K ­ 2FeO4. Appl Surf Sci 346:520–527. https​://doi. org/10.1016/j.apsus​c.2015.04.026 80. Gonzalez-Guerrero AB, Mendoza E, Pellicer E, Alsina F, Fernandez-Sanchez C, Lechuga LM (2008) Discriminating the carboxylic groups from the total acidic sites in oxidized multi-wall carbon nanotubes by means of acid-base titration. Chem Phys Lett 462(4–6):256–259. https​://doi. org/10.1016/j.cplet​t.2008.07.071 81. Riggs JE, Guo ZX, Carroll DL, Sun YP (2000) Strong luminescence of solubilized carbon nanotubes. J Am Chem Soc 122(24):5879–5880 82. Sun YP, Huang WJ, Lin Y, Fu KF, Kitaygorodskiy A, Riddle LA, Yu YJ, Carroll DL (2001) Soluble dendron-functionalized carbon nanotubes: preparation, characterization, and properties. Chem Mater 13(9):2864–2869 83. Chen J, Hamon MA, Hu H, Chen YS, Rao AM, Eklund PC, Haddon RC (1998) Solution properties of single-walled carbon nanotubes. Science 282(5386):95–98 84. Chen J, Rao AM, Lyuksyutov S, Itkis ME, Hamon MA, Hu H, Cohn RW, Eklund PC, Colbert DT, Smalley RE, Haddon RC (2001) Dissolution of full-length single-walled carbon nanotubes. J Phys Chem B 105(13):2525–2528 85. Liu J, Rinzler AG, Dai H, Hafner JH, Bradley RK, Boul PJ, Lu A, Iverson T, Shelimov K, Huffman CB (1998) Fullerene pipes. Science 280(5367):1253–1256 86. Bourlinos AB, Georgakilas V, Tzitzios V, Boukos N, Herrera R, Giannelis ER (2006) Functionalized carbon nanotubes: synthesis of meltable and amphiphilic derivatives. Small 2(10):1188–1191 87. Samori C, Sainz R, Menard-Moyon C, Toma FM, Venturelli E, Singh P, Ballestri M, Prato M, Bianco A (2010) Potentiometric titration as a straightforward method to assess the number of functional groups on shortened carbon nanotubes. Carbon 48(9):2447–2454 88. Calle D, Negri V, Munuera C, Mateos L, Touriño IL, Viñegla PR, Ramírez MO, García-Hernández M, Cerdán S, Ballesteros P (2018) Magnetic anisotropy of functionalized multi-walled carbon nanotube suspensions. Carbon 131:229–237. https​://doi.org/10.1016/j.carbo​n.2018.01.104

Reprinted from the journal

209

13

Topics in Current Chemistry (2020) 378:15 89. Isaac KM, Sabaraya IV, Ghousifam N, Das D, Pekkanen AM, Romanovicz DK, Long TE, Saleh NB, Rylander MN (2018) Functionalization of single-walled carbon nanohorns for simultaneous fluorescence imaging and cisplatin delivery in  vitro. Carbon 138:309–318. https​://doi. org/10.1016/j.carbo​n.2018.06.020 90. Chen RJ, Zhang YG, Wang DW, Dai HJ (2001) Noncovalent sidewall functionalization of singlewalled carbon nanotubes for protein immobilization. J Am Chem Soc 123(16):3838–3839 91. Barry NPE, Therrien B (2016) Pyrene: the guest of honor, chapter 13. In: Sadjadi S (ed) Organic nanoreactors. Academic Press, Boston, pp 421–461. https​://doi.org/10.1016/B978-0-12-80171​ 3-5.00013​-6 92. Bilalis P, Katsigiannopoulos D, Avgeropoulos A, Sakellariou G (2014) Non-covalent functionalization of carbon nanotubes with polymers. RSC Adv 4(6):2911–2934. https​://doi.org/10.1039/ c3ra4​4906h​ 93. Zhu J, Yudasaka M, Zhang MF, Kasuya D, Iijima S (2003) Surface modification approach to the patterned assembly of single-walled carbon nanomaterials. Nano Lett 3(9):1239–1243 94. Zhang T, Ge Y, Wang X, Chen J, Huang X, Liao Y (2017) Polymeric ruthenium porphyrin-functionalized carbon nanotubes and graphene for levulinic ester transformations into γ-valerolactone and pyrrolidone derivatives. ACS Omega 2(7):3228–3240. https​://doi.org/10.1021/acsom​ ega.7b004​27 95. Li H, Zhou B, Lin Y, Gu L, Wang W, Fernando KAS, Kumar S, Allard LF, Sun Y-P (2004) Selective interactions of porphyrins with semiconducting single-walled carbon nanotubes. J Am Chem Soc 126(4):1014–1015. https​://doi.org/10.1021/ja037​142o 96. Hu CY, Xu YJ, Duo SW, Zhang RF, Li MS (2009) Non-covalent functionalization of carbon nanotubes with surfactants and polymers. J Chin Chem Soc-Taip 56(2):234–239. https​://doi. org/10.1002/jccs.20090​0033 97. Vardharajula S, Ali SZ, Tiwari PM, Eroğlu E, Vig K, Dennis VA, Singh SR (2012) Functionalized carbon nanotubes: biomedical applications. Int J Nanomed 7:5361–5374. https​://doi.org/10.2147/ IJN.S3583​2 98. Zhang LW, Zeng L, Barron AR, Monteiro-Riviere NA (2007) Biological interactions of functionalized single-wall carbon nanotubes in human epidermal keratinocytes. Int J Toxicol 26(2):103–113. https​://doi.org/10.1080/10915​81070​12251​33 99. Sadegh H, Shahryari-ghoshekandi R (2015) Functionalization of carbon nanotubes and its application in nanomedicine: a review. Nanomed J 2(4):231–248. https​://doi.org/10.7508/nmj.2015.04.001 100. Vaisman L, Wagner HD, Marom G (2006) The role of surfactants in dispersion of carbon nanotubes. Adv Colloid Interface Sci 128:37–46. https​://doi.org/10.1016/j.cis.2006.11.007 101. Moore VC, Strano MS, Haroz EH, Hauge RH, Smalley RE, Schmidt J, Talmon Y (2003) Individually suspended single-walled carbon nanotubes in various surfactants. Nano Lett 3(10):1379–1382. https​://doi.org/10.1021/nl034​524j 102. Niezabitowska E, Smith J, Prestly MR, Akhtar R, von Aulock Felix W, Lavallée Y, Ali-Boucetta H, McDonald TO (2018) Facile production of nanocomposites of carbon nanotubes and polycaprolactone with high aspect ratios with potential applications in drug delivery. RSC Adv 8(30):16444– 16454. https​://doi.org/10.1039/C7RA1​3553J​ 103. Negri V, Cerpa A, Lopez-Larrubia P, Nieto-Charques L, Cerdan S, Ballesteros P (2010) Nanotubular paramagnetic probes as contrast agents for magnetic resonance imaging based on the diffusion tensor. Angew Chem Int Ed 49(10):1813–1815 104. Cerpa A, Kober M, Calle D, Negri V, Gavira JM, Hernanz A, Briones F, Cerdan S, Ballesteros P (2013) Single-walled carbon nanotubes as anisotropic relaxation probes for magnetic resonance imaging. Medchemcomm 4(4):669–672 105. Bharti A, Cheruvally G (2018) Surfactant assisted synthesis of Pt-Pd/MWCNT and evaluation as cathode catalyst for proton exchange membrane fuel cell. Int J Hydrogen Energy 43(31):14729– 14741. https​://doi.org/10.1016/j.ijhyd​ene.2018.06.009 106. Yasujima R, Yasueda K, Horiba T, Komaba S (2018) Multi-enzyme immobilized anodes utilizing maltose fuel for biofuel cell applications. ChemElectroChem 5(16):2271–2278. https​://doi. org/10.1002/celc.20180​0370 107. Martínez-Paz P, Negri V, Esteban-Arranz A, Martínez-Guitarte JL, Ballesteros P, Morales M (2019) Effects at molecular level of multi-walled carbon nanotubes (MWCNT) in Chironomus riparius (DIPTERA) aquatic larvae. Aquat Toxicol 209:42–48. https​://doi.org/10.1016/j.aquat​ ox.2019.01.017

13

210

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15 108. Ge C, Du J, Zhao L, Wang L, Liu Y, Li D, Yang Y, Zhou R, Zhao Y, Chai Z, Chen C (2011) Binding of blood proteins to carbon nanotubes reduces cytotoxicity. Proc Natl Acad Sci 108(41):16968– 16973. https​://doi.org/10.1073/pnas.11052​70108​ 109. Zhao X, Liu R, Chi Z, Teng Y, Qin P (2010) New insights into the behavior of bovine serum albumin adsorbed onto carbon nanotubes: comprehensive spectroscopic studies. J Phys Chem B 114(16):5625–5631. https​://doi.org/10.1021/jp100​903x 110. Yi CQ, Qi SJ, Zhang DW, Yang MS (2010) Covalent conjugation of multi-walled carbon nanotubes with proteins. Methods Mol Biol 625:9–17. https​://doi.org/10.1007/978-1-60761​-579-8_2 111. Niu S-Y, Li Q-Y, Ren R, Hu K-C (2010) DNA/single-walled carbon nanotubes based fluorescence detection of ­Hg2+. Anal Lett 43(15):2432–2439. https​://doi.org/10.1080/00032​71100​37174​55 112. Tran TL, Nguyen TT, Huyen Tran TT, Chu VT, Thinh Tran Q, Tuan Mai A (2017) Detection of influenza A virus using carbon nanotubes field effect transistor based DNA sensor. Phys E 93:83– 86. https​://doi.org/10.1016/j.physe​.2017.05.019 113. Wu Y, Hudson JS, Lu Q, Moore JM, Mount AS, Rao AM, Alexov E, Ke PC (2006) Coating single-walled carbon nanotubes with phospholipids. J Phys Chem B 110(6):2475–2478. https​://doi. org/10.1021/jp057​252c 114. Liu Z, Tabakman SM, Chen Z, Dai H (2009) Preparation of carbon nanotube bioconjugates for biomedical applications. Nat Protoc 4(9):1372–1382. https​://doi.org/10.1038/nprot​.2009.146 115. Zhou XJ, Moran-Mirabal JM, Craighead HG, McEuen PL (2007) Supported lipid bilayer/carbon nanotube hybrids. Nat Nanotechnol 2(3):185–190. https​://doi.org/10.1038/nnano​.2007.34 116. Yadav P, Rastogi V, Kumar Mishra A, Verma A (2014) Carbon nanotube: a versatile carrier for various biomedical applications. Drug Deliv Lett 4(2):156–169 117. Aqel A, Abou El-Nour KMM, Ammar RAA, Al-Warthan A (2012) Carbon nanotubes, science and technology part (I) structure, synthesis and characterisation. Arab J Chem 5(1):1–23. https​://doi. org/10.1016/j.arabj​c.2010.08.022 118. Belin T, Epron F (2005) Characterization methods of carbon nanotubes: a review. Mater Sci Eng B 119(2):105–118. https​://doi.org/10.1016/j.mseb.2005.02.046 119. Rüther MG, Frehill F, O’Brien JE, Minett AI, Blau WJ, Vos JG, in het Panhuis M (2004) Characterization of covalent functionalized carbon nanotubes. J Phys Chem B 108(28):9665–9668. https​ ://doi.org/10.1021/jp040​266i 120. Dresselhaus MS, Jorio A, Hofmann M, Dresselhaus G, Saito R (2010) Perspectives on carbon nanotubes and graphene Raman spectroscopy. Nano Lett 10(3):751–758. https​://doi.org/10.1021/ nl904​286r 121. Jorio A, Fantini C, Dantas MSS, Pimenta MA, Souza Filho AG, Samsonidze GG, Brar VW, Dresselhaus G, Dresselhaus MS, Swan AK, Ünlü MS, Goldberg BB, Saito R (2002) Linewidth of the Raman features of individual single-wall carbon nanotubes. Phys Rev B 66(11):115411. https​://doi. org/10.1103/PhysR​evB.66.11541​1 122. Henrard L, Popov VN, Rubio A (2001) Influence of packing on the vibrational properties of infinite and finite bundles of carbon nanotubes. Phys Rev B 64(20):205403. https​://doi.org/10.1103/ PhysR​evB.64.20540​3 123. Henrard L, Hernández E, Bernier P, Rubio A (1999) van der Waals interaction in nanotube bundles: consequences on vibrational modes. Phys Rev B 60(12):R8521–R8524. https​://doi.org/10.1103/ PhysR​evB.60.R8521​ 124. Wepasnick KA, Smith BA, Bitter JL, Howard Fairbrother D (2010) Chemical and structural characterization of carbon nanotube surfaces. Anal Bioanal Chem 396(3):1003–1014. https​://doi. org/10.1007/s0021​6-009-3332-5 125. Umemura K, Izumi K, Oura S (2016) Probe microscopic studies of DNA molecules on carbon nanotubes. Nanomaterials (Basel). https​://doi.org/10.3390/nano6​10018​0 126. Thostenson ET, Ren ZF, Chou TW (2001) Advances in the science and technology of carbon nanotubes and their composites: a review. Compos Sci Technol 61(13):1899–1912. https​://doi. org/10.1016/S0266​-3538(01)00094​-X 127. Meunier V, Lambin P (2004) Scanning tunnelling microscopy of carbon nanotubes. Philos Trans A Math Phys Eng Sci 362(1823):2187–2203. https​://doi.org/10.1098/rsta.2004.1435 128. Bychko I, Strizhak P (2018) Carbon nanotubes catalytic activity in the ethylene hydrogenation. Fullerenes Nanotubes Carbon Nanostruct 26(12):804–809. https​://doi.org/10.1080/15363​ 83X.2018.15021​76

Reprinted from the journal

211

13

Topics in Current Chemistry (2020) 378:15 129. Okpalugo TIT, Papakonstantinou P, Murphy H, McLaughlin J, Brown NMD (2005) High resolution XPS characterization of chemical functionalised MWCNTs and SWCNTs. Carbon 43(1):153– 161. https​://doi.org/10.1016/j.carbo​n.2004.08.033 130. Peigney A, Laurent C, Flahaut E, Bacsa RR, Rousset A (2001) Specific surface area of carbon nanotubes and bundles of carbon nanotubes. Carbon 39(4):507–514. https​://doi.org/10.1016/S0008​ -6223(00)00155​-X 131. Berber S, Kwon YK, Tomanek D (2000) Unusually high thermal conductivity of carbon nanotubes. Phys Rev Lett 84(20):4613–4616 132. Salvetat J-P, Bonard J-M, Thomson NH, Kulik AJ, Forró L, Benoit W, Zuppiroli L (1999) Mechanical properties of carbon nanotubes. Appl Phys A 69(3):255–260. https​://doi.org/10.1007/s0033​ 90050​999 133. Chłopek J, Czajkowska B, Szaraniec B, Frackowiak E, Szostak K, Béguin F (2006) In vitro studies of carbon nanotubes biocompatibility. Carbon 44(6):1106–1111. https​://doi.org/10.1016/j.carbo​ n.2005.11.022 134. Fernandes LF, Bruch GE, Massensini AR, Frezard F (2018) Recent advances in the therapeutic and diagnostic use of liposomes and carbon nanomaterials in ischemic stroke. Front Neurosci 12:453. https​://doi.org/10.3389/fnins​.2018.00453​ 135. Tilmaciu CM, Morris MC (2015) Carbon nanotube biosensors. Front Chem 3:59. https​://doi. org/10.3389/fchem​.2015.00059​ 136. Pasinszki T, Krebsz M, Tung TT, Losic D (2017) Carbon nanomaterial based biosensors for noninvasive detection of cancer and disease biomarkers for clinical diagnosis. Sensors (Basel). https​:// doi.org/10.3390/s1708​1919 137. Zhou Y, Fang Y, Ramasamy RP (2019) Non-covalent functionalization of carbon nanotubes for electrochemical biosensor development. Sensors (Basel). https​://doi.org/10.3390/s1902​0392 138. Song CK, Oh E, Kang MS, Shin BS, Han SY, Jung M, Lee ES, Yoon SY, Sung MM, Ng WB, Cho NJ, Lee H (2018) Fluorescence-based immunosensor using three-dimensional CNT network structure for sensitive and reproducible detection of oral squamous cell carcinoma biomarker. Anal Chim Acta 1027:101–108. https​://doi.org/10.1016/j.aca.2018.04.025 139. Qian P, Qin Y, Lyu Y, Li Y, Wang L, Wang S, Liu Y (2019) A hierarchical cobalt/carbon nanotube hybrid nanocomplex-based ratiometric fluorescent nanosensor for ultrasensitive detection of hydrogen peroxide and glucose in human serum. Anal Bioanal Chem 411(8):1517–1524. https​:// doi.org/10.1007/s0021​6-019-01573​-z 140. Holzinger M, Baur J, Haddad R, Wang X, Cosnier S (2011) Multiple functionalization of single-walled carbon nanotubes by dip coating. Chem Commun 47(8):2450–2452. https​://doi. org/10.1039/C0CC0​3928D​ 141. Tang X, Bansaruntip S, Nakayama N, Yenilmez E, Chang YL, Wang Q (2006) Carbon nanotube DNA sensor and sensing mechanism. Nano Lett 6(8):1632–1636. https​://doi.org/10.1021/nl060​ 613v 142. Lee J, Morita M, Takemura K, Park EY (2018) A multi-functional gold/iron-oxide nanoparticleCNT hybrid nanomaterial as virus DNA sensing platform. Biosens Bioelectron 102:425–431. https​ ://doi.org/10.1016/j.bios.2017.11.052 143. Chen Y, Guo S, Zhao M, Zhang P, Xin Z, Tao J, Bai L (2018) Amperometric DNA biosensor for Mycobacterium tuberculosis detection using flower-like carbon nanotubes-polyaniline nanohybrid and enzyme-assisted signal amplification strategy. Biosens Bioelectron 119:215–220. https​://doi. org/10.1016/j.bios.2018.08.023 144. Nouri M, Meshginqalam B, Sahihazar MM, Sheydaie Pour Dizaji R, Ahmadi MT, Ismail R (2018) Experimental and theoretical investigation of sensing parameters in carbon nanotube-based DNA sensor. IET Nanobiotechnol 12(8):1125–1129. https​://doi.org/10.1049/iet-nbt.2018.5068 145. Sun Y, Peng Z, Li H, Wang Z, Mu Y, Zhang G, Chen S, Liu S, Wang G, Liu C, Sun L, Man B, Yang C (2019) Suspended CNT-Based FET sensor for ultrasensitive and label-free detection of DNA hybridization. Biosens Bioelectron 137:255–262. https​://doi.org/10.1016/j.bios.2019.04.054 146. Gong H, Peng R, Liu Z (2013) Carbon nanotubes for biomedical imaging: the recent advances. Adv Drug Deliv Rev 65(15):1951–1963. https​://doi.org/10.1016/j.addr.2013.10.002 147. Kuznik N, Tomczyk MM (2016) Multiwalled carbon nanotube hybrids as MRI contrast agents. Beilstein J Nanotechnol 7:1086–1103. https​://doi.org/10.3762/bjnan​o.7.102 148. Gao Y (2018) Carbon nano-allotrope/magnetic nanoparticle hybrid nanomaterials as T2 contrast agents for magnetic resonance imaging applications. J Funct Biomater. https​://doi.org/10.3390/ jfb90​10016​

13

212

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15



149. Ania Servant IJ, Bussy Cyrill, Fabbro Chiara, da Ros Tatiana, Pach Elzbieta, Belen Ballesteros MP, Nicolay Klaas, Kostarelos Kostas (2016) Gadolinium-functionalised multi-walled carbon nanotubes as a T1 contrast agent for MRI cell labelling and tracking. Carbon 97:8 150. Mehri-Kakavand G, Hasanzadeh H, Jahanbakhsh R, Abdollahi M, Nasr R, Bitarafan-Rajabi A, Jadidi M, Darbandi-Azar A, Emadi A (2019) Gdn (3 +)@CNTs-PEG versus Gadovist(R): in vitro assay. Oman Med J 34(2):147–155. https​://doi.org/10.5001/omj.2019.27 151. Moghaddam SE, Hernandez-Rivera M, Zaibaq NG, Ajala A, da Graca Cabreira-Hansen M, Mowlazadeh-Haghighi S, Willerson JT, Perin EC, Muthupillai R, Wilson LJ (2018) A new high-performance gadonanotube-polymer hybrid material for stem cell labeling and tracking by MRI. Contrast Media Mol Imaging 2018:2853736. https​://doi.org/10.1155/2018/28537​36 152. Antaris AL, Yaghi OK, Hong G, Diao S, Zhang B, Yang J, Chew L, Dai H (2015) Single chirality (6,4) single-walled carbon nanotubes for fluorescence imaging with silicon detectors. Small 11(47):6325–6330. https​://doi.org/10.1002/smll.20150​1530 153. Yudasaka M, Yomogida Y, Zhang M, Tanaka T, Nakahara M, Kobayashi N, Okamatsu-Ogura Y, Machida K, Ishihara K, Saeki K, Kataura H (2017) Near-infrared photoluminescent carbon nanotubes for imaging of brown fat. Sci Rep UK 7:44760. https​://doi.org/10.1038/srep4​4760 154. Kim J-W, Galanzha EI, Shashkov EV, Moon H-M, Zharov VP (2009) Golden carbon nanotubes as multimodal photoacoustic and photothermal high-contrast molecular agents. Nat Nanotechnol 4(10):688–694. https​://doi.org/10.1038/nnano​.2009.231 155. Liang C, Diao S, Wang C, Gong H, Liu T, Hong G, Shi X, Dai H, Liu Z (2014) Tumor metastasis inhibition by imaging-guided photothermal therapy with single-walled carbon nanotubes. Adv Mater 26(32):5646–5652. https​://doi.org/10.1002/adma.20140​1825 156. McDevitt MR, Chattopadhyay D, Jaggi JS, Finn RD, Zanzonico PB, Villa C, Rey D, Mendenhall J, Batt CA, Njardarson JT, Scheinberg DA (2007) PET imaging of soluble yttrium-86-labeled carbon nanotubes in mice. PLoS One 2(9):e907–e907. https​://doi.org/10.1371/journ​al.pone.00009​07 157. Georgin D, Czarny B, Botquin M, Mayne-L’Hermite M, Pinault M, Bouchet-Fabre B, Carriere M, Poncy J-L, Chau Q, Maximilien R, Dive V, Taran F (2009) Preparation of 14C-labeled multiwalled carbon nanotubes for biodistribution investigations. J Am Chem Soc 131(41):14658–14659. https​ ://doi.org/10.1021/ja906​319z 158. Wang H, Wang J, Deng X, Sun H, Shi Z, Gu Z, Liu Y, Zhaoc Y (2004) Biodistribution of carbon single-wall carbon nanotubes in mice. J Nanosci Nanotechnol 4(8):1019–1024. https​://doi. org/10.1166/jnn.2004.146 159. Guo J, Zhang X, Li Q, Li W (2007) Biodistribution of functionalized multiwall carbon nanotubes in mice. Nucl Med Biol 34(5):579–583. https​://doi.org/10.1016/j.nucme​dbio.2007.03.003 160. Liu Z, Cai W, He L, Nakayama N, Chen K, Sun X, Chen X, Dai H (2007) In vivo biodistribution and highly efficient tumour targeting of carbon nanotubes in mice. Nat Nanotechnol 2(1):47–52 161. Al-Jamal KT, Nunes A, Methven L, Ali-Boucetta H, Li S, Toma FM, Herrero MA, Al-Jamal WT, ten Eikelder HMM, Foster J, Mather S, Prato M, Bianco A, Kostarelos K (2012) Degree of chemical functionalization of carbon nanotubes determines tissue distribution and excretion profile. Angew Chem Int Ed 51(26):6389–6393. https​://doi.org/10.1002/anie.20120​1991 162. Raphey VR, Henna TK, Nivitha KP, Mufeedha P, Sabu C, Pramod K (2019) Advanced biomedical applications of carbon nanotube. Mater Sci Eng C Mater Biol Appl 100:616–630. https​://doi. org/10.1016/j.msec.2019.03.043 163. Panwar N, Soehartono AM, Chan KK, Zeng S, Xu G, Qu J, Coquet P, Yong KT, Chen X (2019) Nanocarbons for biology and medicine: sensing, imaging, and drug delivery. Chem Rev. https​:// doi.org/10.1021/acs.chemr​ev.9b000​99 164. Mohajeri M, Behnam B, Sahebkar A (2018) Biomedical applications of carbon nanomaterials: drug and gene delivery potentials. J Cell Physiol 234(1):298–319. https​://doi.org/10.1002/ jcp.26899​ 165. Bianco A, Kostarelos K, Prato M (2005) Applications of carbon nanotubes in drug delivery. Curr Opin Chem Biol 9(6):674–679 166. Rawal S, Patel MM (2019) Threatening cancer with nanoparticle aided combination oncotherapy. J Control Release 301:76–109. https​://doi.org/10.1016/j.jconr​el.2019.03.015 167. Liu D, Zhang Q, Wang J, Fan L, Zhu W, Cai D (2019) Hyaluronic acid-coated single-walled carbon nanotubes loaded with doxorubicin for the treatment of breast cancer. Pharmazie 74(2):83–90. https​://doi.org/10.1691/ph.2019.8152

Reprinted from the journal

213

13

Topics in Current Chemistry (2020) 378:15 168. Kam NWS, Liu Z, Dai H (2005) Functionalization of carbon nanotubes via cleavable disulfide bonds for efficient intracellular delivery of siRNA and potent gene silencing. J Am Chem Soc 127(36):12492–12493. https​://doi.org/10.1021/ja053​962k 169. Taghavi S, Nia AH, Abnous K, Ramezani M (2017) Polyethylenimine-functionalized carbon nanotubes tagged with AS1411 aptamer for combination gene and drug delivery into human gastric cancer cells. Int J Pharm 516(1):301–312. https​://doi.org/10.1016/j.ijpha​rm.2016.11.027 170. Guo C, Al-Jamal WT, Toma FM, Bianco A, Prato M, Al-Jamal KT, Kostarelos K (2015) Design of cationic multiwalled carbon nanotubes as efficient siRNA vectors for lung cancer xenograft eradication. Bioconjug Chem 26(7):1370–1379. https​://doi.org/10.1021/acs.bioco​njche​m.5b002​49 171. Foldvari M, Bagonluri M (2008) Carbon nanotubes as functional excipients for nanomedicines: II. Drug delivery and biocompatibility issues. Nanomedicine Nanotechnol Biol Med 4(3):183–200. https​://doi.org/10.1016/j.nano.2008.04.003 172. Son KH, Hong JH, Lee JW (2016) Carbon nanotubes as cancer therapeutic carriers and mediators. Int J Nanomed 11:5163–5185. https​://doi.org/10.2147/IJN.S1126​60 173. Guo Q, Shen XT, Li YY, Xu SQ (2017) Carbon nanotubes-based drug delivery to cancer and brain. J Huazhong Univ Sci Technol Med Sci 37(5):635–641. https​://doi.org/10.1007/s1159​6-017-1783-z 174. Moon HK, Lee SH, Choi HC (2009) In  vivo near-infrared mediated tumor destruction by photothermal effect of carbon nanotubes. ACS Nano 3(11):3707–3713. https​://doi.org/10.1021/nn900​ 904h 175. Needa AV, Carole D, Patrick M, Paul H, Robert EH, Joel S, Ricardo PS, Daniel ER, Roger GH (2018) Phosphatidylserine targeted single-walled carbon nanotubes for photothermal ablation of bladder cancer. Nanotechnology 29(3):035101 176. Shiue RJ, Gao Y, Tan C, Peng C, Zheng J, Efetov DK, Kim YD, Hone J, Englund D (2019) Thermal radiation control from hot graphene electrons coupled to a photonic crystal nanocavity. Nat Commun 10(1):109. https​://doi.org/10.1038/s4146​7-018-08047​-3 177. Xu Y, Shan Y, Cong H, Shen Y, Yu B (2018) Advanced carbon-based nanoplatforms combining drug delivery and thermal therapy for cancer treatment. Curr Pharm Des 24(34):4060–4076. https​ ://doi.org/10.2174/13816​12825​66618​11201​60959​ 178. Wang L, Shi J, Zhang H, Li H, Gao Y, Wang Z, Wang H, Li L, Zhang C, Chen C, Zhang Z, Zhang Y (2013) Synergistic anticancer effect of RNAi and photothermal therapy mediated by functionalized single-walled carbon nanotubes. Biomaterials 34(1):262–274. https​://doi.org/10.1016/j.bioma​ teria​ls.2012.09.037 179. Kafa H, Wang JT-W, Rubio N, Venner K, Anderson G, Pach E, Ballesteros B, Preston JE, Abbott NJ, Al-Jamal KT (2015) The interaction of carbon nanotubes with an in  vitro blood-brain barrier model and mouse brain in vivo. Biomaterials 53:437–452. https​://doi.org/10.1016/j.bioma​teria​ ls.2015.02.083 180. Kafa H, Wang JT-W, Rubio N, Klippstein R, Costa PM, Hassan HAFM, Sosabowski JK, Bansal SS, Preston JE, Abbott NJ, Al-Jamal KT (2016) Translocation of LRP1 targeted carbon nanotubes of different diameters across the blood–brain barrier in vitro and in vivo. J Control Release 225:217–229. https​://doi.org/10.1016/j.jconr​el.2016.01.031 181. Wang JTW, Rubio N, Kafa H, Venturelli E, Fabbro C, Ménard-Moyon C, Da Ros T, Sosabowski JK, Lawson AD, Robinson MK, Prato M, Bianco A, Festy F, Preston JE, Kostarelos K, Al-Jamal KT (2016) Kinetics of functionalised carbon nanotube distribution in mouse brain after systemic injection: spatial to ultra-structural analyses. J Control Release 224:22–32. https​://doi. org/10.1016/j.jconr​el.2015.12.039 182. Pedro Miguel Costa JT-WW, Morfin Jean-François, Khanum Tamanna, To Wan, Sosabowski Jane, Tóth Eva, Al-Jamal Khuloud T (2018) Functionalised carbon nanotubes enhance brain delivery of amyloid-targeting Pittsburgh compound B (PiB)-derived ligands. Nanotheranostics 2(2):168–183. https​://doi.org/10.7150/ntno.23125​ 183. Lohan S, Raza K, Mehta SK, Bhatti GK, Saini S, Singh B (2017) Anti-Alzheimer’s potential of berberine using surface decorated multi-walled carbon nanotubes: a preclinical evidence. Int J Pharm 530(1):263–278. https​://doi.org/10.1016/j.ijpha​rm.2017.07.080 184. Hassanzadeh P, Arbabi E, Atyabi F, Dinarvand R (2017) Nerve growth factor-carbon nanotube complex exerts prolonged protective effects in an in  vitro model of ischemic stroke. Life Sci 179:15–22. https​://doi.org/10.1016/j.lfs.2016.11.029 185. Fiorito S, Russier J, Salemme A, Soligo M, Manni L, Krasnowska E, Bonnamy S, Flahaut E, Serafino A, Togna GI, Marlier LNJL, Togna AR (2018) Switching on microglia with electro-conductive multi walled carbon nanotubes. Carbon 129:572–584. https​://doi.org/10.1016/j.carbo​n.2017.12.069

13

214

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15 186. Lee HJ, Park J, Yoon OJ, Kim HW, Lee DY, Kim DH, Lee WB, Lee N-E, Bonventre JV, Kim SS (2011) Amine-modified single-walled carbon nanotubes protect neurons from injury in a rat stroke model. Nat Nanotechnol 6(2):121–125. https​://doi.org/10.1038/nnano​.2010.281 187. Cellot GLM, Scaini D, Rauti R, Bosi S, Prato M, Gandolfi S, Ballerini L (2017) Successful regrowth of retinal neurons when cultured interfaced to carbon nanotube platforms. J Biomed Nanotechnol 13(5):559–567. https​://doi.org/10.1166/jbn.2017.2364 188. Salehi M, Naseri-Nosar M, Ebrahimi-Barough S, Nourani M, Khojasteh A, Hamidieh A-A, Amani A, Farzamfar S, Ai J (2018) Sciatic nerve regeneration by transplantation of Schwann cells via erythropoietin controlled-releasing polylactic acid/multiwalled carbon nanotubes/gelatin nanofibrils neural guidance conduit. J Biomed Mater Res B Appl Biomater 106(4):1463–1476. https​:// doi.org/10.1002/jbm.b.33952​ 189. Xue X, Yang J-Y, He Y, Wang L-R, Liu P, Yu L-S, Bi G-H, Zhu M-M, Liu Y-Y, Xiang R-W, Yang X-T, Fan X-Y, Wang X-M, Qi J, Zhang H-J, Wei T, Cui W, Ge G-L, Xi Z-X, Wu C-F, Liang X-J (2016) Aggregated single-walled carbon nanotubes attenuate the behavioural and neurochemical effects of methamphetamine in mice. Nat Nanotechnol 11(7):613–620. https​://doi.org/10.1038/ nnano​.2016.23 190. Saliev T (2019) The advances in biomedical applications of carbon nanotubes. C 5:22. https​://doi. org/10.3390/c5020​029 191. Mashal A, Sitharaman B, Li X, Avti PK, Sahakian AV, Booske JH, Hagness SC (2010) Toward carbon-nanotube-based theranostic agents for microwave detection and treatment of breast cancer: enhanced dielectric and heating response of tissue-mimicking materials. IEEE Trans Biomed Eng 57(8):1831–1834. https​://doi.org/10.1109/TBME.2010.20425​97 192. Al Faraj A, Shaik AS, Ratemi E, Halwani R (2016) Combination of drug-conjugated SWCNT nanocarriers for efficient therapy of cancer stem cells in a breast cancer animal model. J Control Release 225:240–251. https​://doi.org/10.1016/j.jconr​el.2016.01.053 193. Zhang M, Wang W, Wu F, Yuan P, Chi C, Zhou N (2017) Magnetic and fluorescent carbon nanotubes for dual modal imaging and photothermal and chemo-therapy of cancer cells in living mice. Carbon 123:70–83. https​://doi.org/10.1016/j.carbo​n.2017.07.032 194. Hou L, Yang X, Ren J, Wang Y, Zhang H, Feng Q, Shi Y, Shan X, Yuan Y, Zhang Z (2016) A novel redox-sensitive system based on single-walled carbon nanotubes for chemo-photothermal therapy and magnetic resonance imaging. Int J Nanomed 11:607–624. https​://doi.org/10.2147/IJN. S9847​6 195. Zhang M, Wang Wentao, Wu Fan, Yuan Ping, Chi Cheng, Zhou Ninglin (2017) Magnetic and fluorescent carbon nanotubes for dual modal imaging and photothermal and chemo-therapy of cancer cells in living mice. Carbon 123:14. https​://doi.org/10.1016/j.carbo​n.2017.07.032 196. Wang X, Wang C, Cheng L, Lee ST, Liu Z (2012) Noble metal coated single-walled carbon nanotubes for applications in surface enhanced Raman scattering imaging and photothermal therapy. J Am Chem Soc 134(17):7414–7422. https​://doi.org/10.1021/ja300​140c 197. Zhao H, Chao Y, Liu J, Huang J, Pan J, Guo W, Wu J, Sheng M, Yang K, Wang J, Liu Z (2016) Polydopamine coated single-walled carbon nanotubes as a versatile platform with radionuclide labeling for multimodal tumor imaging and therapy. Theranostics 6(11):1833–1843. https​://doi. org/10.7150/thno.16047​ 198. Lovat V, Pantarotto D, Lagostena L, Cacciari B, Grandolfo M, Righi M, Spalluto G, Prato M, Ballerini L (2005) Carbon nanotube substrates boost neuronal electrical signaling. Nano Lett 5(6):1107–1110. https​://doi.org/10.1021/nl050​637m 199. Mazzatenta A, Giugliano M, Campidelli S, Gambazzi L, Businaro L, Markram H, Prato M, Ballerini L (2007) Interfacing neurons with carbon nanotubes: electrical signal transfer and synaptic stimulation in cultured brain circuits. J Neurosci 27(26):6931–6936. https​://doi.org/10.1523/ JNEUR​OSCI.1051-07.2007 200. Ren J, Xu Q, Chen X, Li W, Guo K, Zhao Y, Wang Q, Zhang Z, Peng H, Li Y-G (2017) Superaligned carbon nanotubes guide oriented cell growth and promote electrophysiological homogeneity for synthetic cardiac tissues. Adv Mater 29(44):1702713. https​://doi.org/10.1002/adma.20170​ 2713 201. Silva E, Vasconcellos LMRd, Rodrigues BVM, dos Santos DM, Campana-Filho SP, Marciano FR, Webster TJ, Lobo AO (2017) PDLLA honeycomb-like scaffolds with a high loading of superhydrophilic graphene/multi-walled carbon nanotubes promote osteoblast in vitro functions and guided in vivo bone regeneration. Mater Sci Eng C 73:31–39. https​://doi.org/10.1016/j.msec.2016.11.075

Reprinted from the journal

215

13

Topics in Current Chemistry (2020) 378:15 202. Gutiérrez-Hernández JM, Escobar-García DM, Escalante A, Flores H, González FJ, Gatenholm P, Toriz G (2017) In vitro evaluation of osteoblastic cells on bacterial cellulose modified with multiwalled carbon nanotubes as scaffold for bone regeneration. Mater Sci Eng C 75:445–453. https​:// doi.org/10.1016/j.msec.2017.02.074 203. Lima MD, Hussain MW, Spinks GM, Naficy S, Hagenasr D, Bykova JS, Tolly D, Baughman RH (2015) Efficient, absorption-powered artificial muscles based on carbon nanotube hybrid yarns. Small 11(26):3113–3118. https​://doi.org/10.1002/smll.20150​0424 204. Liu ZF, Fang S, Moura FA, Ding JN, Jiang N, Di J, Zhang M, Lepro X, Galvao DS, Haines CS, Yuan NY, Yin SG, Lee DW, Wang R, Wang HY, Lv W, Dong C, Zhang RC, Chen MJ, Yin Q, Chong YT, Zhang R, Wang X, Lima MD, Ovalle-Robles R, Qian D, Lu H, Baughman RH (2015) Stretchy electronics. Hierarchically buckled sheath-core fibers for superelastic electronics, sensors, and muscles. Science 349(6246):400–404. https​://doi.org/10.1126/scien​ce.aaa79​52 205. Vashist A, Kaushik A, Vashist A, Sagar V, Ghosal A, Gupta YK, Ahmad S, Nair M (2018) Advances in carbon nanotubes-hydrogel hybrids in nanomedicine for therapeutics. Adv Healthc Mater 7(9):e1701213. https​://doi.org/10.1002/adhm.20170​1213 206. Shin SR, Jung SM, Zalabany M, Kim K, Zorlutuna P, Kim SB, Nikkhah M, Khabiry M, Azize M, Kong J, Wan KT, Palacios T, Dokmeci MR, Bae H, Tang XS, Khademhosseini A (2013) Carbon-nanotube-embedded hydrogel sheets for engineering cardiac constructs and bioactuators. ACS Nano 7(3):2369–2380. https​://doi.org/10.1021/nn305​559j 207. Joddar B, Garcia E, Casas A, Stewart CM (2016) Development of functionalized multi-walled carbon-nanotube-based alginate hydrogels for enabling biomimetic technologies. Sci Rep 6:32456. https​://doi.org/10.1038/srep3​2456 208. Sun H, Zhou J, Huang Z, Qu L, Lin N, Liang C, Dai R, Tang L, Tian F (2017) Carbon nanotubeincorporated collagen hydrogels improve cell alignment and the performance of cardiac constructs. Int J Nanomed 12:3109–3120. https​://doi.org/10.2147/IJN.S1280​30 209. Roshanbinfar K, Mohammadi Z, Sheikh-Mahdi Mesgar A, Dehghan MM, Oommen OP, Hilborn J, Engel FB (2019) Carbon nanotube doped pericardial matrix-derived electroconductive biohybrid hydrogel for cardiac tissue engineering. Biomater Sci. https​://doi.org/10.1039/c9bm0​0434c​ 210. Lee JR, Ryu S, Kim S, Kim BS (2015) Behaviors of stem cells on carbon nanotube. Biomater Res 19:3. https​://doi.org/10.1186/s4082​4-014-0024-9 211. Jan E, Kotov NA (2007) Successful differentiation of mouse neural stem cells on layer-by-layer assembled single-walled carbon nanotube composite. Nano Lett 7(5):1123–1128. https​://doi. org/10.1021/nl062​0132 212. Chao TI, Xiang S, Chen CS, Chin WC, Nelson AJ, Wang C, Lu J (2009) Carbon nanotubes promote neuron differentiation from human embryonic stem cells. Biochem Biophys Res Commun 384(4):426–430. https​://doi.org/10.1016/j.bbrc.2009.04.157 213. Namgung S, Kim T, Baik KY, Lee M, Nam JM, Hong S (2011) Fibronectin-carbon-nanotube hybrid nanostructures for controlled cell growth. Small 7(1):56–61. https​://doi.org/10.1002/ smll.20100​1513 214. Lichtenstein MP, Carretero NM, Perez E, Pulido-Salgado M, Moral-Vico J, Sola C, Casan-Pastor N, Sunol C (2018) Biosafety assessment of conducting nanostructured materials by using cocultures of neurons and astrocytes. Neurotoxicology 68:115–125. https​://doi.org/10.1016/j.neuro​ .2018.07.010 215. Ramanathan M, Patil M, Epur R, Yun Y, Shanov V, Schulz M, Heineman WR, Datta MK, Kumta PN (2016) Gold-coated carbon nanotube electrode arrays: immunosensors for impedimetric detection of bone biomarkers. Biosens Bioelectron 77:580–588. https​://doi.org/10.1016/j. bios.2015.10.014 216. Prato M, Kostarelos K, Bianco A (2008) Functionalized carbon nanotubes in drug design and discovery. Acc Chem Res 41(1):60–68. https​://doi.org/10.1021/ar700​089b 217. Mazzaglia A, Scala A, Sortino G, Zagami R, Zhu Y, Sciortino MT, Pennisi R, Pizzo MM, Neri G, Grassi G, Piperno A (2018) Intracellular trafficking and therapeutic outcome of multiwalled carbon nanotubes modified with cyclodextrins and polyethylenimine. Colloids Surf B Biointerfaces 163:55–63. https​://doi.org/10.1016/j.colsu​rfb.2017.12.028

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

13

216

Reprinted from the journal

Topics in Current Chemistry (2020) 378:15

Affiliations Viviana Negri1   · Jesús Pacheco‑Torres2   · Daniel Calle3   · Pilar López‑Larrubia4  1

Departamento de Biotecnología y Farmacia, Facultad de Ciencias Biomédicas, Universidad Europea de Madrid, Villaviciosa de Odón, Spain

2

Division of Cancer Imaging Research, The Russell H. Morgan Department of Radiology and Radiological Science, The Johns Hopkins University School of Medicine, Baltimore, MD, USA

3

Laboratorio de Imagen Médica, Hospital Universitario Gregorio Marañón, c/Dr. Esquerdo 56, 28007 Madrid, Spain

4

Instituto de Investigaciones Biomédicas “Alberto Sols”, CSIC-UAM, c/Arturo Duperier 4, 28029 Madrid, Spain





Reprinted from the journal

217

13

Topics in Current Chemistry (2020) 378:8 https://doi.org/10.1007/s41061-019-0273-0 REVIEW

Bioconjugated Plasmonic Nanoparticles for Enhanced Skin Penetration David Alba‑Molina1 · Juan J. Giner‑Casares1 · Manuel Cano1  Received: 7 November 2019 / Accepted: 3 December 2019 / Published online: 16 December 2019 © Springer Nature Switzerland AG 2019

Abstract Plasmonic nanoparticles (NPs) are one of the most promising and studied inorganic nanomaterials for different biomedical applications. Plasmonic NPs have excellent biocompatibility, long-term stability against physical and chemical degradation, relevant optical properties, well-known synthesis methods and tuneable surface functionalities. Herein, we review recently reported bioconjugated plasmonic NPs using different chemical approaches and loading cargoes (such as drugs, genes, and proteins) for enhancement of transdermal delivery across biological tissues. The main aim is to understand the interaction of the complex skin structure with biomimetic plasmonic NPs. This knowledge is not only important in enhancing transdermal delivery of pharmaceutical formulations but also for controlling undesired skin penetration of industrial products, such as cosmetics, sunscreen formulations and any other mass-usage consumable that contains plasmonic NPs. Keywords  Plasmonic nanoparticles · Transdermal drug delivery · Bioconjugated nanomaterials · Skin penetration · Gold and silver nanoparticles

Chapter 7 was originally published as Alba‑Molina, D., Giner‑Casares, J. J. & Cano, M. Topics in Current Chemistry (2020) 378: 8. https://doi.org/10.1007/s41061-019-0273-0. * Juan J. Giner‑Casares [email protected] * Manuel Cano [email protected] 1



Department of Physical Chemistry and Applied Thermodynamics, Institute of Nanochemistry (IUNAN), University of Córdoba, Campus Universitario de Rabanales, Ed. Marie Curie, 14014 Córdoba, Spain

Reprinted from the journal

219

13

Vol.:(0123456789)



Topics in Current Chemistry (2020) 378:8

1 Introduction This review is focused on plasmonic nanoparticles (NPs) from among the whole range of inorganic NPs. Plasmonic NPs are composed of colloidal particles of noble metals that present a characteristic surface plasmon resonance (SPR) band. The SPR band is attributed to the electric field of incident light, which induces coherent oscillation of conduction band electrons of the positively charged metallic core (Fig. 1). The optical and electronic properties of plasmonic NPs can be tuned easily by changing their size, shape, and surface chemistry [1–3]. Among the different plasmonic NPs, this work is focused mainly on Au and Ag-based NPs, which are considered as non-allergenic compounds that should not induce cytotoxicity, and are therefore highly attractive for biomedical applications [4, 5]. In addition to the inherent characteristic properties of nanometric materials, such as small size and high surface-area-to-volume ratio, the surface chemistry of AuNPs can be modified easily through covalent Au–S bonds [6, 7]. These features have made AuNPs one of the most widely used nanomaterials in both technological and biomedical applications, such as electronics [8, 9], catalysis [10, 11], sensory probes [12, 13], plasmonic photothermal therapy [14, 15], targeted drug delivery

Fig. 1  Schematic representation of surface plasmon resonance (SPR) band in A spherical and B rodshaped plasmonic nanoparticles (NPs)

13

220

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8

carrier [16, 17], contrast agents in X-ray imaging and computed tomography for cancer diagnosis [18, 19], to name only a few. AgNPs display an inherent antimicrobial capacity against bacteria, viruses, and other eukaryotic microorganisms [20, 21]. Indeed, silver compounds have been used historically to prevent microbial growth in many fields, such as in wound care and in products such as odor-reducing clothing, acne creams, and face masks [22, 23]. Topical administration of compounds presents several advantages compared with other routes such as oral, nasal, and intravenous administration. Transdermal delivery overcomes first-pass hepatic metabolism and can reduce harmful side-effects [24–26]. However, skin is a complex multilayer structure (composed mainly of epidermis, dermis, and hypodermis) and is a highly impermeable barrier to most molecules on the basis of particle size, water-solubility and surface charge [24]. In addition, hair follicles have also been explored as a more permeable transport channel for transdermal drug delivery [27–29]. An excellent review of the influence of size on the penetration of both metal and non-metal NPs through skin was published by Larese Filon et al. [30]. Naked NPs do not overcome skin barriers efficiently, as reported by NúñezLozano et  al. [31] In other words, non-functionalized NPs constitute poor transdermal drug delivery systems [32, 33]. Therefore, many efforts have been made to enhance the skin penetration of NPs through purposely designed chemical functionalization with biomolecules and bioinspired polymers to form biomimetic NPs providing additional abilities in skin penetration. Herein, the latest advances in the development of different approaches to fabricate bioconjugated plasmonic NPs for enhanced skin penetration are presented. This review is organized as follows: first, the influence on the skin penetration of key parameters of non-bioconjugated plasmonic NPs is discussed, and, second, some of the more recently reported bioconjugated plasmonic NPs for enhanced skin penetration are presented.

2 Influence of Key‑Parameters on the Skin Penetration of Non‑bioconjugated Plasmonic NPs The ability of plasmonic NPs to penetrate skin can be tuned by controlling their core-composition, particle-size and -shape, surface charge, water solubility and functionalization strategy, including both capping-ligands and delivered cargoes, i.e., drugs, genes, and proteins. Relevant examples of the most prominent studies in each of these parameters are mentioned below. 2.1 Core Composition (Noble Metal Type) The literature indicates a significant difference in terms of skin penetration behavior between organic and inorganics NPs. There are also significant differences in skin internalization depending on the chemical composition (i.e. T ­ iO2, ­SiO2, ZnO, FeO, CdSe, Pd, etc.) within the latter group, as previously reported by Larese Filon et  al. [30]. In the case of plasmonic NPs, in  vitro penetration of AgNPs through

Reprinted from the journal

221

13

Topics in Current Chemistry (2020) 378:8

intact human skin samples has been reported for several authors (with particle size in a narrow range between 19 and 25 nm), suggesting that ion release is the most feasible penetration mechanism [34, 35]. Whilst, for AuNPs, penetration through intact human skin samples (size ~ 12 nm) has been demonstrated by several authors, although the penetration process is unclear, although the ions release mechanism could be discounted. The absence of Au ions in physiological solutions of AuNPs has been reported, indicating higher long-term colloidal stability than AgNPs [30, 36]. This high stability could explain the absence or low cytotoxicity of this type of plasmonic NPs. 2.2 Particle Size Effects Although the literature reports contradictory results on this issue, skin penetration of NPs is considered a size-dependent process [30]. Note that the stability of the NP coating should be taken into account because it plays a key role in the interactions between skin and both NP core (see section on  Influence of the Capping-Ligand, Including Surface-Chemistry and -Charge), and the NPs (i.e., steric stabilization) by avoiding particle aggregation when they come into contact with the stratum corneum and constituent cells of the skin, mainly keratinocytes. A representative study showing that AuNPs penetrate through intact skin samples in a size-dependent manner was reported by Sonavane et al. [37], who analyzed the penetration of 15, 102 and 198 nm citrate-capped AuNPs, with spherical shapes and surface negative charge, through rat-skin and rat-intestine using Franz diffusion cells. The smallest AuNPs showed higher permeation than the larger particles. In a similar study with rats, Raju and co-workers reported that 22  nm citrate-capped AuNPs showed higher penetration than 105 and 186  nm particles across the thick stratum corneum of the plantar rat skin [38]. A key aspect of the experimental design for assessing the skin penetration of plasmonic NPs is the in vitro or in vivo model used to perform the analysis. Indeed, the choice of a biologically relevant and realistic model for studying the biological effect of NPs is a comparatively unexplored field [39]. Although rat, mouse and rabbits have been used extensively for these penetration studies, pig skin is probably the most similar animal model to the human skin [40]. Pig and human skin are structurally very similar in thickness and dermal-epidermal thickness ratio. Hair follicles and blood vessel patterns in the skin are also similar. In addition, the thickness of the human skin varies considerably as a function of the body region, gender and age, among other factors. For this reason, the results obtained are quite different depending on the in vivo model used, and even depending on the part of the human body used. An interesting study on porcine skin using AgNPs was reported by Samberg and co-workers [22], who evaluated the in  vitro and in  vivo toxicity of eight different commercial AgNPs supplied by nanoComposix (San Diego, CA), such as unwashed/uncoated (diameter of 20, 50, and 80 nm), washed/uncoated (20, 50, and 80  nm), and carbon-coated AgNPs (25 and 35  nm). They observed that the toxicity of AgNPs in human embryonic kidney (HEKs) was influenced significantly by residual contaminants in their supernatant, and that AgNPs themselves may not be

13

222

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8

responsible for the observed increase in cell mortality. The degradation of AgNPs within the cell was also considered as a source of reactive oxygen species that would be damaging to the cell machinery and DNA [41]. Obviously, human skin would be ideal to perform in vitro skin penetration analysis. In 2012, Liu et co-workers [42] investigated the in vitro penetration and metabolic effects of 10, 30 and 60 nm citrate-capped AuNPs within viable excised human abdominal skin after 24-h exposure. Using multiphoton tomograph-fluorescence lifetime imaging microscopy, these authors observed penetration only into the stratum corneum, without significant penetration into the lower layers. They demonstrated that viable human skin resists permeation of small NPs, which had been previously reported to penetrate deeply in other animal skin models. Note that a remarkable attempt to categorize the size range with its skin penetration ability was carried out by Larese Filon et al. [30] They suggested that: (1) NPs smaller than 4 nm can both penetrate and permeate intact skin, (2) NPs in the size range between 4 and 20  nm can permeate both intact and damaged skin, (3) NPs with diameter between 21 and 45 nm can penetrate and permeate damaged skin, and (4) NPs higher than 45 nm can neither penetrate nor permeate the skin. They also considered the hydrodynamic diameter of the NPs, which is an important parameter of the colloidal particles that can be affected greatly both in terms of ligand-coating and the electrolyte composition of the colloidal solution [43]. 2.3 Shape‑Effect Despite the enormous literature reporting protocols for the synthesis of plasmonic NPs with different morphologies, such as rod, triangle, bipyramid, star, cube and others, [44–46], as well as the important effects on the resulting NP flow characteristics, with altered cell membrane interactions, macrophage uptake and circulating lifetimes [47–49], research papers investigating the influence of the NPs shape on skin penetration are rare (Fig.  2). This fact can be attributed mainly to the resulting higher size (e.g. > 45  nm) and to the lower resulting long-term stability of the anisotropic NPs (i.e. non-spherical geometries). Only comparative studies between spherical and rod-shape AuNPs can be found in the literature. A relevant study was reported by Fernandes et al. [50], with their culture experiments in mouse and human skin samples showing that the percentage of PEG-capped Au-nanorods (with an aspect ratio 2.8 ± 0.5) found in all samples was higher than that obtained in similar PEG-capped spherical 15 ± 1 nm AuNPs. These results were obtained for both positively and negatively surface charged NPs, suggesting the great influence of NPshape on penetration capacity. In the case of AgNPs, Tak and co-workers reported a skin penetration study of differently shaped NPs using both in  vitro and in  vivo models [51]. They used spherical, rod-shape and triangular AgNPs with similar hydrodynamic diameter (~ 50 nm) and zeta-potential (+ 30 mV) to perform in vitro analysis on ultra-thin mouse skin section by the Franz cell system, and in vivo analysis on hairless mice. In agreement with previous results for AuNPs, they showed that rod-shaped AgNPs presented a higher permeability index than spherical and triangular AgNPs. They concluded that different shapes of AgNPs may exhibit diverse Reprinted from the journal

223

13



Topics in Current Chemistry (2020) 378:8

Fig. 2  Schematic representation of A plasmonic NPs with different geometries and B human skin layers

antimicrobial activities and skin penetration capabilities depending upon their active metallic facets. 2.4 Influence of the Capping‑Ligand, Including Surface‑Chemistry and ‑Charge Other important parameters that greatly affects the skin penetration rate of NPs are the surface charge and chemistry, which are associated directly with the stabilizing ligand composition and the type of metal–ligand bonding interaction. Figure 3 summarizes the most frequently found capping ligands, including the type of interaction with the metallic core. Citrate is the main ligand for the synthesis of plasmonic NPs [52, 53], and is the most widely reported. Although the resulting water-soluble NPs present good long-term stability, with a negatively charged surface due to their carboxylate groups, the NPs tended to aggregate easily in contact with the skin (especially for the smallest sizes) due to their relatively weak metal–ligand interactions [7]. Labouta and co-workers [54] showed that water-soluble spherical-shape 15 nm citrate-capped AuNPs tended to aggregate on the superficial stratum corneum. In this study, the penetration rate of hydrophilic 15  nm citrate-capped AuNPs versus hydrophobic 6  nm dodecanethiol-capped AuNPs was also compared, showing that non-water-soluble particles penetrated through the stratum corneum and into viable epidermal layers of human skin. This enhanced skin penetration using dodecanethiol as capping ligand could be attributed to: (1) stronger Au–S binding interaction, which avoided aggregation of NPs in contact with the stratum corneum; and (2) the solubility in organic solvent (i.e., toluene), which could facilitate interaction with cell membranes. A related study from the same research group compared the different penetration rates through human skin of two water-soluble AuNPs (15 nm citrate- and 6 nm lecithin-capped) against two toluene-solved AuNPs (6 nm dodecanethiol- and cetrimide-capped) [55]. They concluded that the vehicle (toluene-versus-water) had a minimal effect on skin penetration of AuNPs.

13

224

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8

Fig. 3  Representative stabilizing ligand for plasmonic NPs

Not only should the bond strength of the metal–ligand interaction be considered but also the coating thickness and the degree of surface coverage. Previous studies have demonstrated that polymeric capping ligands, such as poly(N-vinyl-2-pyrrolidone) (PVP), are well suited to stabilizing the surface of AuNPs for skin penetration, despite the weak metal–ligand interactions via pyrrolidone groups [56]. Huang and co-workers [56] showed that water-soluble 5  nm PVP-capped AuNPs with a spherical shape were mice skin permeable. They attributed this effect to the nanobio interaction with skin lipids, and the consequent induction of transient and reversible openings on the stratum corneum. In addition, they highlighted that co-administration of these PVP-capped AuNPs with protein drugs could enhance transdermal drug delivery. Reprinted from the journal

225

13



Topics in Current Chemistry (2020) 378:8

Additionally, the capping ligand confers a given surface charge on NPs, which greatly affects the resulting skin penetration rate. For instance, positively charged drug carriers, such as dendrimers and liposomes, are well-known to induce greater drug delivery in the skin [57, 58]. In this context, Fernandes et al. [50] showed that positive charged PEG-capped AuNPs were found in the skin at levels from  2- to 6-fold higher than in their negative counterparts. These results were obtained both for spherical- and for rod-shaped NPs, and were in agreement with the enhanced skin permeation of cationic liposomes, which was attributed to the “Donnan exclusion effect” and to the better interaction of cationic particles with negatively charged skin cells [59]. Furthermore, in this study, they also showed that AuNPs functionalized with cell penetrating peptides (CPPs) TAT and R7 were found in the skin in larger quantities than PEGylated AuNPs, demonstrating that bioconjugation greatly enhances skin penetration rate [50]. In an alternative strategy, Lee and co-workers investigated the influence of the surface charge of Au-nanorods on skin penetration using a layer‐by‐layer (LbL) polyelectrolyte coating technique [60]. They observed that negative charged CTAB/ PSS-capped Au-nanorods penetrated more rapidly through the skin than the positive ones (CTAB- and CTAB/PSS/PDADMAC-capped). For this, three different multi-layer coated Au-nanorods with a particle size of 18 × 40  nm were used: two positively charged, CTAB- and CTAB/PSS/PDADMAC-capped, and one negatively charged, CTAB/PSS-capped. These surprising results were attributed to both aggregation of the positively charged Au-nanorods on the stratum corneum and the adsorption of proteins released from the dermis layer to the surface of Au-nanorods. In line with these results, Mahmoud et al. [61] observed that positively charged Aunanorods aggregated extensively upon exposure to human skin compared to their negatively and neutrally charged counterparts. They attributed these findings to the adsorption of proteins released from the dermis layer to the surface of Au-nanorods. In this latter study, they prepared 49.5 × 12.0 nm Au-nanorods capped with four different surface ligands: cetyltrimethylammonium (CTAB), polyacrylic acid (PAA), poly(allylamine hydrochloride) (PAH), and methoxy-polyethylene glycol-thiol (m-PEG-SH). Conversely, Hao et al. [62] also investigated the influence of the surface charge on the skin penetration of spherical AuNPs using human reconstructed 3D Episkin model. In this study, three different surfaces charged 5  nm-AuNPs capped with citrate (negative), PVP (neutral), and CTAB (positive) were tested. They observed that, although all AuNPs induced the phase change of lipid lamella and passed through the epidermis, positively charged AuNPs exhibited the most efficient skin penetration through both the paracellular routes and the transcellular pathway when compared to neutrally or negatively charged NPs. An interesting alternative for stabilizing NP surfaces is PEGylation, which is an approach commonly used for improving the drug and gene delivery efficiency of NP-based systems to target cells and tissues [63]. Hsiao and co-workers [64] employed this approach to investigate the positive effects of polyethylene glycol (HS-PEG-COOH) and HS-PEG-oleylamine (OAm) functionalization on the skin permeation of spherical 10 nm AuNPs. Using an in vivo rat model, they showed that PEG- and PEG-OAm-functionalized AuNPs were able to overcome the skin barrier and deposit in the deeper subcutaneous adipose tissue. Moreover, the follicular

13

226

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8



deposition of AuNPs increased 2-fold after PEG-OAm functionalization, demonstrating a preferential accumulation mediated by the stabilizing ligand. Mahmoud et  al. [65] evaluated the preferential accumulation of Au-nanorods into abdominal human skin hair follicles. To this end, they prepared 11.4 × 46.6  nm Au-nanorods with five different surface chemistries (i.e., neutral, anionic, cationic, and hydrophobic), such as CTAB, PAA, methoxy-polyethylene glycol-thiol (m-PEG-SH), PEG-Cystamine, and polystyrene (PS). They observed that the lipophilic properties of sebum-rich hair follicles enhanced the accumulation of hydrophobic PS-Aunanorods into hair follicles, while neutral m-PEG-S-Au-nanorods were distributed into all skin compartments, especially the dermis, which exhibits hydrophilic characteristics. In addition, both charged Au-nanorods showed a negligible percentage of penetration into any of the skin compartments.

3 Bioconjugated Plasmonic NPs for Transdermal Delivery of Different Cargoes A seminal study on the development of bioconjugated plasmonic NPs for enhancing the skin penetration of different cargoes was reported in 2010 by Huang et al. [56]. They demonstrated significant enhancement of the transdermal delivery of protein-drugs by co-administration with 5  nm PVP-capped AuNPs. This fact was attributed to the nano-bio interaction with skin lipids, which allowed a reversible openings of the stratum corneum. Thus, this work provided a simple and efficient NP-mediated method for overcoming the skin barrier for percutaneous protein drug delivery. Labala et al. [66] reported the first bioconjugated plasmonic NPs for iontophoretic transdermal delivery of imatinib mesylate to treat melanoma, using an LbL assembly approach. This LbL polymer capped AuNP contained PVP and polyethylene imine (PEI), was subsequentially coated with anionic poly(styrenesulfonate) (PSS) and cationic PEI for drug loading. The resulting bioconjugated nanosystem showed an average particle size and a zeta-potential of 98 ± 4 nm and + 32 ± 1 mV, respectively, and a shift in the SPR wavelength from 518 to 530  nm. The in  vitro skin penetration studies were performed on excised porcine ear, and demonstrated that iontophoresis application enhanced the skin penetration of imatinib mesylate loaded AuNP by 6.2-fold compared with passive application. Bessar et  al. [67] prepared water-soluble sodium 3-mercapto1-propansulfonate(3MPS)-capped AuNPs, which were loaded with methotrexate (MTX) via electrostatic adsorption. The resulting Au-3MPS@MTX conjugate showed an average size and a zeta-potential of ~5 nm and −32 ± 1 mV, respectively. It was then administrated topically on C57BL/6 mouse normal skin in order to assess its absorption behaviour. In vitro and in vivo studies showed that MTX-conjugated AuNPs were much more efficient than MTX alone, suggesting this nanosystem as a potential candidate for topical treatment of psoriasis (see Fig. 4A). An interesting application of bioconjugated plasmonic NPs for enhanced transdermal gene delivery was the reported by Niu and co-workers [69]. In order to facilitate the skin penetration of plasmid DNA (i.e., pDNA encoding miRNA-221 inhibitor -Mi221-) deep into melanoma tissues, these authors synthesized 20–25 nm Reprinted from the journal

227

13



Topics in Current Chemistry (2020) 378:8

Fig. 4  A Electron microscopy images of normal human keratinocytes. a Control, b cells treated with bioconjugated plasmonic NPs [67]. Copyright Elsevier, 2016. B Wounds treated with non-coated and bioconjugated plasmonic NPs, showing the absence of granulation tissue [68]. Copyright Elsevier, 2018

conjugated AuNPs containing a cell-penetrating TAT peptide and the cationic PEI that could compact the pDNAs into cationic nanocomplexes (i.e., zeta-potential ~ +35  mV). They demonstrated that the resulting plasmonic bioconjugates can penetrate through the intact stratum corneum without any additional physical enhancement method. This study proposed a novel topical gene therapy strategy for skin cancer with great priority to reverse both the progression and metastasis of advanced melanoma. Chen et  al. [70] fabricated a biocongugate plasmonic NP for the transdermal delivery of vascular endothelial growth factor (VEGF) in wound repair. To that end, they performed the bioconjugation of AuNP-PEG-COOH with VEGF through carbodiimide bonds, obtaining a negative surface charged nanosystem, whose absorption capability was evaluated by a mouse skin model. After treatment, they observed not only the presence of VEGF in the dermis but also its effect for promoting angiogenesis, demonstrating that, in this case, the binding of protein biological factors to AuNPs could preserve the activity of the protein. Another study was reported by Safwat et  al., who fabricated AuNPs capped with benzalkonium chloride and with PEI for enhanced loading and skin permeability of 5-fluorouracil (i.e., 5-FU/BC-AuNPs and 5-FU/PEI-AuNPs, respectively) [71]. They performed ex  vivo permeability studies of different 5-FU preparations using mice skin, demonstrating that the permeability of 5-FU was significantly higher for drug-loaded AuNPs compared with the other tested 5-FU samples. This same research group also

13

228

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8



prepared 5-FU loaded through ionic interactions onto CTAB capped AuNPs, and the resulting nanocomposite was incorporated into gel and cream bases to evaluate its permeability both ex vivo in mice dorsal skin and in vivo in A431 tumor-bearing mice [72]. They observed that the nano-formulation provided around 2-fold higher permeability through mice skin compared with free 5-FU gel and cream formulations, and achieved 6.8- and 18.4-fold lower tumour volume than the untreated control with the gel- and the cream-based nano-formulation, respectively. On the other hand, Boca et  al. [73] performed the first preliminar study to evaluate the potential use under dermatological conditions of Ruxolitinib-conjugated 15  nm AuNPs as alternative for treating alopecia. Using in  vitro preclinical setting, they showed that AuNPs@TWEEN-20@Ruxolitinib inhibited the proliferation of fibroblasts by inhibiting JAK2 protein, suggesting it as a potential strategy to treat alopecia. Another novel and alternative administration strategy of bioconjugated plasmonic NPs was proposed by Anirudhan et al. [74], who fabricated a nanocomposite film containing methacrylate-stitched β-cyclodextrin embedded with AuNPs and hydrophobic titanium nanotube (TNT) and tested the transdermal delivery of ibuprofen through in vitro rat skin. They showed that the resulting film exhibited an improved drug-delivery performance, which was attributed to synergistic action of AuNP and hydrophobic TNT. They proposed this nanocomposite film as an alternative skin permeation strategy for transdermal drug delivery. Similarly, this same research group proposed a polyelectrolyte membrane fabricated with guar gum, poly(vinyl alcohol) and a nanogold-nanocellulose composite for the topical administration of diltiazem hydrochloride. In vivo use of this film on human skin was analyzed, suggesting its potential use for transdermal drug delivery [75]. Pan et al. [68] explored the effects on wound healing of keratinocyte growth factor (KGF) cross-linked to AuNPs. Using an animal full-thickness wound model, they showed that KGF-AuNPs were more favorable to wound healing than bare AuNPs or KGF, thus proposing KGF-AuNPs as a promising wound healing drug for clinical application, see Fig. 4B. Crisan et al. [76] evaluated the impact on psoriatic inflammation of AgNPs and AuNPs complexed with Cornus mas (i.e., polyphenols-rich extracts) by using an in vitro model based on pro-inflammatory macrophages. The results obtained from all the performed in  vitro analysis suggested that these bioconjugated plasmonic NPs provide an efficient tool for modern psoriasis therapy, circumventing immunosuppression-related side effects of biologicals. In another study, Wang et al. [77] fabricated an antimicrobial peptide (LL37) grafted ultra-small AuNPs (AuNPs@LL37, ~7 nm), which was combined with proangiogenic (VEGF) plasmids to analyze its potential use for the topical treatment of diabetic wounds with or without bacterial infection. The resulting bioconjugate (AuNPs@LL37/pDNAs) combined the advantages of cationic surface charged NPs that condense DNA with those of antibacterial peptides and enhance the cellular and nucleus entry to achieve high gene delivery efficiency. AuNPs@LL37/pDNAs were shown to greatly improve the gene transfection efficiency in keratinocytes compared with pristine AuNPs/pDNAs, exhibiting a similar expression to Lipo2000/pDNAs (a well-known highly efficient gene transfection agent), whilst displaying higher antibacterial ability. Thus, this bioconjugated plasmonic NPs were suggested as a suitable strategy for treating chronic diabetic wounds. Reprinted from the journal

229

13



Topics in Current Chemistry (2020) 378:8

More recently, Fratoddi et al. [78] analyzed the effects of AuNPs functionalized with 3-mercapto-1-propansulfonate (AuNPs-3MPS) and loading MTX topically administered in vitro on a skin model and in vivo on an imiquimod-induced psoriasis-like mice model. The showed that treatment with this system was able to induce a reduction in keratinocyte hyperproliferation, epidermal thickness and also the volume of inflammatory infiltrate in the in vivo model used. Hernández-Martínez et al. [79] synthetized and evaluated a nanocomposite of AuNPs functionalized with calreticulin. Using in  vitro and in  vivo wound healing mice models of diabetes, they assessed the ability of the nanocomposite to promote proliferation and migration. Their results confirmed the utility of this bioconjugated plasmonic NPs (AuNPscalreticulin) as potential treatment for wound healing of diabetic ulcers. In the case of AgNPs-based bioconjugates, Mandal and co-workers [80] fabricated a nanocomposite hydrogel comprised of in  situ formed Ag nanowires (AgNWs) deposited with chemically cross-linked carboxymethyl cellulose (CMC), which demonstrated superior efficacy as a transdermal anticancer drug-curcumin carrier. This plasmonic bioconjugate had the capacity to encapsulate both hydrophobic/hydrophilic transdermal drugs. In vitro experiments suggested that the presence of AgNWs on cross-linked CMC enhanced both the penetration power of nanocomposite hydrogel and drug release in a sustained manner. Whilst ex vivo rat skin permeation analysis confirmed that drug delivery through the nanocomposite hydrogel was permeable through the rat skin in controlled fashion, efficiently killing the MG 63 cancer cells. Table 1 summarizes the bioconjugated plasmonic NPs cited in this review, indicating both the loaded active molecule and the potential application.

4 Conclusions Bioconjugated plasmonic NPs are a promising approach for topical administration of different cargos for several diseases. The excellent biocompatibility and readily adjustable physical and chemical features of plasmonic NPs are highly attractive options for purposefully designed nanomaterials aimed at biomedical applications. Several examples have been presented herein, illustrating the wide range of cargoes and functionalization strategies that might be included when designing a bioconjugated plasmonic NPs. Different physical and chemical parameters should be taken into account when analyzing the effect of plasmonic NPs on human skin. While chemical routes for obtaining on-demand plasmonic NPs are relatively wellestablished, and a large number of simple and reproducible experimental protocols are available, the main frontier for mass usage is still a correct assessment of the toxicity of the NPs. Providing a relevant model for human skin, the experimental conditions for studying location and local concentration of plasmonic NPs differ greatly from those found in synthesis laboratories. Therefore, this fruitful field of research requires more efforts to fully understand the penetration mechanisms of these bioconjugated plasmonic NPs, enabling a decrease in associated toxicity and potential long-term environmental impacts. In view of the latest contribution to the field, we speculate that a reliable framework will be available in the short term, enabling a

13

230

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8 Table 1  Representative cargoes in through bioconjugated plasmonic nanoparticles (NPs) for skin disease treatments Active molecule

Cargo type

Activity/application

NP Au

References

Imatinib mesylate

Drug

Anticancer

Methotrexate (MTX)

Drug

Anti-inflammatory (psoriasis) Au

[67, 78]

[66]

microRNA mir-221

Gene

Tumour suppressor (melanoma)

Au

[69]

Vascular endothelial growth factor (VEGF)

Protein

Wound repair

Au

[70]

Ibuprofen

Drug

Anti-inflammatory

Au

[74]

Diltiazem hydrochloride

Drug

Vasodilator

Au

[75]

5-fluorouracil (5-FU)

Drug

Anticancer

Au

[71, 72]

Ruxolitinib

Drug

Anti-alopecia

Au

[73]

Keratinocyte growth factor (KGF)

Protein

Wound repair

Au

[68]

Polyphenols-rich extracts (Cornus mas)

Natural extract

Anti-inflammatory (psoriasis) Au & Ag [76]

Antimicrobial peptide LL-37 and pDNA: Pro-angiogenic (VEGF) plasmids

Protein and gene Diabetic wound healing

Au

[77]

Calreticulin

Protein

Diabetic wound healing

Au

[79]

Curcumin

Natural extract

Anticancer

Ag

[80]

second wave of research of directed synthesis and application of plasmonic NPs under in vivo conditions. Acknowledgements  We apologize to authors whose work could not be included in this review due to space restrictions. Support from the Ministry of Science, Innovation and Universities of Spain is acknowledged through the MANA project CTQ2017-83961-R  and JEANS project  CTQ2017-92264-EXP. J. J. G.-C. acknowledges the Ministry of Science, Innovation and Universities of Spain for a “Ramon y Cajal” contract (#RyC-2014-14956). M. C. thanks the “Plan Propio de Investigación” from the Universidad de Córdoba (UCO) and the “Programa Operativo de fondos FEDER Andalucía” for its financial support through both postdoctoral contracts (Modality 5.2.A).

References 1. Huang X, Jain PK, El-Sayed IH, El-Sayed MA (2007) Gold nanoparticles: interesting optical properties and recent applications in cancer diagnostics and therapy. Nanomedicine 2:681–693. https​:// doi.org/10.2217/17435​889.2.5.681 2. Giner-Casares JJ, Henriksen-Lacey M, Coronado-Puchau M, Liz-Marzán LM (2016) Inorganic nanoparticles for biomedicine: where materials scientists meet medical research. Mater Today 19:19– 28. https​://doi.org/10.1016/j.matto​d.2015.07.004 3. Ogarev VA, Rudoi VM, Dement’eva OV (2018) Gold nanoparticles: synthesis, optical properties, and application. Inorg Mater Appl Res 9:134–140. https​://doi.org/10.1134/S2075​11331​80101​97 4. Shukla R, Bansal V, Chaudhary M et  al (2005) Biocompatibility of gold nanoparticles and their endocytotic fate inside the cellular compartment: a microscopic overview. Langmuir 21:10644– 10654. https​://doi.org/10.1021/la051​3712 Reprinted from the journal

231

13



Topics in Current Chemistry (2020) 378:8

5. Burduşel A-C, Gherasim O, Grumezescu AM et al (2018) Biomedical applications of silver nanoparticles: an up-to-date overview. Nanomaterials 8:681. https​://doi.org/10.3390/nano8​09068​1 6. Woehrle GH, Brown LO, Hutchison JE (2005) Thiol-functionalized, 1.5-nm gold nanoparticles through ligand exchange reactions: scope and mechanism of ligand exchange. J Am Chem Soc 127:2172–2183. https​://doi.org/10.1021/ja045​7718 7. Alba-Molina D, Puente Santiago AR, Giner-Casares JJ et  al (2019) Tailoring the ORR and HER electrocatalytic performances of gold nanoparticles through metal–ligand interfaces. J Mater Chem A 7:20425–20434. https​://doi.org/10.1039/C9TA0​5492H​ 8. Berry V, Saraf RF (2005) Self-assembly of nanoparticles on live bacterium: an avenue to fabricate electronic devices. Angew Chemie Int Ed 44:6668–6673. https​://doi.org/10.1002/anie.20050​1711 9. Chen Y-S, Hong M-Y, Huang GS (2012) A protein transistor made of an antibody molecule and two gold nanoparticles. Nat Nanotechnol 7:197–203. https​://doi.org/10.1038/nnano​.2012.7 10. Alba-Molina D, Rodriguez-Padron D, Puente Santiago ARR et al (2018) Mimicking the bioelectrocatalytic function of recombinant CotA laccase via electrostatically self-assembled nanobioconjugates. Nanoscale. https​://doi.org/10.1039/C8NR0​6001K​ 11. Alba-Molina D, Puente Santiago AR, Giner-Casares JJ et al (2019) Citrate-stabilized gold nanoparticles as high-performance electrocatalysts: the role of size in the electroreduction of oxygen. J Phys Chem C 123:9807–9812. https​://doi.org/10.1021/acs.jpcc.9b002​49 12. Upadhyayula VKK (2012) Functionalized gold nanoparticle supported sensory mechanisms applied in detection of chemical and biological threat agents: a review. Anal Chim Acta 715:1–18. https​:// doi.org/10.1016/j.aca.2011.12.008 13. Draz MS, Shafiee H (2018) Applications of gold nanoparticles in virus detection. Theranostics 8:1985–2017. https​://doi.org/10.7150/thno.23856​ 14. Huang X, El-Sayed MA (2010) Gold nanoparticles: optical properties and implementations in cancer diagnosis and photothermal therapy. J Adv Res 1:13–28. https​://doi.org/10.1016/j. jare.2010.02.002 15. Vines JB, Yoon J-H, Ryu N-E et  al (2019) Gold nanoparticles for photothermal cancer therapy. Front Chem 7:167. https​://doi.org/10.3389/fchem​.2019.00167​ 16. Dreaden EC, Austin LA, Mackey MA, El-Sayed MA (2012) Size matters: gold nanoparticles in targeted cancer drug delivery. Ther Deliv 3:457–478. https​://doi.org/10.4155/tde.12.21 17. Farooq MU, Novosad V, Rozhkova EA et al (2018) Gold nanoparticles-enabled efficient dual delivery of anticancer therapeutics to HeLa cells. Sci Rep 8:2907. https​://doi.org/10.1038/s4159​8-01821331​-y 18. Cole LE, Ross RD, Tilley JM et al (2015) Gold nanoparticles as contrast agents in x-ray imaging and computed tomography. Nanomedicine 10:321–341. https​://doi.org/10.2217/nnm.14.171 19. Mahan MM, Doiron AL (2018) Gold nanoparticles as X-ray, CT, and multimodal imaging contrast agents: formulation, targeting, and methodology. J Nanomater 2018:1–15. https​://doi. org/10.1155/2018/58372​76 20. Gong P, Li H, He X et al (2007) Preparation and antibacterial activity of F ­ e3O4 @Ag nanoparticles. Nanotechnology 18:285604. https​://doi.org/10.1088/0957-4484/18/28/28560​4 21. Durán N, Durán M, de Jesus MB et  al (2016) Silver nanoparticles: a new view on mechanistic aspects on antimicrobial activity. Nanomed Nanotechnol Biol Med 12:789–799. https​://doi. org/10.1016/j.nano.2015.11.016 22. Samberg ME, Oldenburg SJ, Monteiro-Riviere NA (2010) Evaluation of silver nanoparticle toxicity in skin in vivo and keratinocytes in vitro. Environ Health Perspect 118:407–413. https​://doi. org/10.1289/ehp.09013​98 23. Brandt O, Mildner M, Egger AE et al (2012) Nanoscalic silver possesses broad-spectrum antimicrobial activities and exhibits fewer toxicological side effects than silver sulfadiazine. Nanomed Nanotechnol Biol Med 8:478–488. https​://doi.org/10.1016/j.nano.2011.07.005 24. Pegoraro C, MacNeil S, Battaglia G (2012) Transdermal drug delivery: from micro to nano. Nanoscale 4:1881. https​://doi.org/10.1039/c2nr1​1606e​ 2 5. Amjadi M, Mostaghaci B, Sitti M (2017) Recent advances in skin penetration enhancers for transdermal gene and drug delivery. Curr Gene Ther 17:139–146. https​://doi.org/10.2174/15665​23217​ 66617​05101​51540​ 2 6. Prausnitz MR, Mitragotri S, Langer R (2004) Current status and future potential of transdermal drug delivery. Nat Rev Drug Discov 3:115–124. https​://doi.org/10.1038/nrd13​04 27. Patzelt A, Lademann J (2013) Drug delivery to hair follicles. Expert Opin Drug Deliv 10:787–797. https​://doi.org/10.1517/17425​247.2013.77603​8

13

232

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8



28. Lauterbach A, Müller-Goymann CC (2015) Applications and limitations of lipid nanoparticles in dermal and transdermal drug delivery via the follicular route. Eur J Pharm Biopharm 97:152–163. https​://doi.org/10.1016/j.ejpb.2015.06.020 29. Vogt A, Wischke C, Neffe AT et  al (2016) Nanocarriers for drug delivery into and through the skin—do existing technologies match clinical challenges? J Control Release 242:3–15. https​://doi. org/10.1016/j.jconr​el.2016.07.027 30. Larese Filon F, Mauro M, Adami G et  al (2015) Nanoparticles skin absorption: new aspects for a safety profile evaluation. Regul Toxicol Pharmacol 72:310–322. https​://doi.org/10.1016/j.yrtph​ .2015.05.005 31. Núñez-Lozano R, Cano M, Pimentel B, de la Cueva-Méndez G (2015) ‘Smartening’ anticancer therapeutic nanosystems using biomolecules. Curr Opin Biotechnol 35:135–140. https​://doi. org/10.1016/j.copbi​o.2015.07.005 32. Campbell CSJ, Contreras-Rojas LR, Delgado-Charro MB, Guy RH (2012) Objective assessment of nanoparticle disposition in mammalian skin after topical exposure. J Control Release 162:201–207. https​://doi.org/10.1016/j.jconr​el.2012.06.024 33. Zhou Y, Damasceno PF, Somashekar BS et al (2018) Unusual multiscale mechanics of biomimetic nanoparticle hydrogels. Nat Commun 9:181. https​://doi.org/10.1038/s4146​7-017-02579​-w 34. Larese FF, D’Agostin F, Crosera M et  al (2009) Human skin penetration of silver nanoparticles through intact and damaged skin. Toxicology 255:33–37. https​://doi.org/10.1016/j.tox.2008.09.025 35. Bianco C, Adami G, Crosera M et al (2014) Silver percutaneous absorption after exposure to silver nanoparticles: a comparison study of three human skin graft samples used for clinical applications. Burns 40:1390–1396. https​://doi.org/10.1016/j.burns​.2014.02.003 36. Larese Filon F, Crosera M, Adami G et  al (2011) Human skin penetration of gold nanoparticles through intact and damaged skin. Nanotoxicology 5:493–501. https​://doi.org/10.3109/17435​ 390.2010.55142​8 37. Sonavane G, Tomoda K, Sano A et  al (2008) In  vitro permeation of gold nanoparticles through rat skin and rat intestine: effect of particle size. Colloids Surf B Biointerfaces 65:1–10. https​://doi. org/10.1016/j.colsu​rfb.2008.02.013 38. Raju G, Katiyar N, Vadukumpully S, Shankarappa SA (2018) Penetration of gold nanoparti cles across the stratum corneum layer of thick-Skin. J Dermatol Sci 89:146–154. https​://doi. org/10.1016/j.jderm​sci.2017.11.001 39. Rivera-Gil P, Jimenez De Aberasturi D, Wulf V et  al (2013) The challenge to relate the physicochemical properties of colloidal nanoparticles to their cytotoxicity. Acc Chem Res 46:743–749. https​://doi.org/10.1021/ar300​039j 40. Avci P, Sadasivam M, Gupta A et  al (2013) Animal models of skin disease for drug discovery. Expert Opin Drug Discov 8:331–355. https​://doi.org/10.1517/17460​441.2013.76120​2 41. Arora S, Jain J, Rajwade JM, Paknikar KM (2008) Cellular responses induced by silver nanoparticles: in vitro studies. Toxicol Lett 179:93–100. https​://doi.org/10.1016/j.toxle​t.2008.04.009 42. Liu DC, Raphael AP, Sundh D et al (2012) The human stratum corneum prevents small gold nanoparticle penetration and their potential toxic metabolic consequences. J Nanomater 2012:1–8. https​ ://doi.org/10.1155/2012/72170​6 43. Cano M, Núñez-Lozano R, Lumbreras R et al (2017) Partial PEGylation of superparamagnetic iron oxide nanoparticles thinly coated with amine-silane as a source of ultrastable tunable nanosystems for biomedical applications. Nanoscale 9:812–822. https​://doi.org/10.1039/C6NR0​7462F​ 44. Grzelczak M, Pérez-Juste J, Mulvaney P, Liz-Marzán LM (2008) Shape control in gold nanoparticle synthesis. Chem Soc Rev 37:1783. https​://doi.org/10.1039/b7114​90g 45. Hühn J, Carrillo-Carrion C, Soliman MG et al (2017) Selected standard protocols for the synthesis, phase transfer, and characterization of inorganic colloidal nanoparticles. Chem Mater 29:399–461. https​://doi.org/10.1021/acs.chemm​ater.6b047​38 46. Sánchez-Iglesias A, Winckelmans N, Altantzis T et al (2017) High-yield seeded growth of monodisperse pentatwinned gold nanoparticles through thermally induced seed twinning. J Am Chem Soc 139:107–110. https​://doi.org/10.1021/jacs.6b121​43 47. Toy R, Peiris PM, Ghaghada KB, Karathanasis E (2014) Shaping cancer nanomedicine: the effect of particle shape on the in  vivo journey of nanoparticles. Nanomedicine 9:121–134. https​://doi. org/10.2217/nnm.13.191 48. Blanco E, Shen H, Ferrari M (2015) Principles of nanoparticle design for overcoming biological barriers to drug delivery. Nat Biotechnol 33:941–951. https​://doi.org/10.1038/nbt.3330

Reprinted from the journal

233

13

Topics in Current Chemistry (2020) 378:8 49. Wang W, Gaus K, Tilley RD, Gooding JJ (2019) The impact of nanoparticle shape on cellular internalisation and transport: what do the different analysis methods tell us? Mater Horizons 6:1538–1547. https​://doi.org/10.1039/C9MH0​0664H​ 5 0. Fernandes R, Smyth NR, Muskens OL et al (2015) Interactions of skin with gold nanoparticles of different surface charge, shape, and functionality. Small 11:713–721. https​://doi.org/10.1002/ smll.20140​1913 51. Tak YK, Pal S, Naoghare PK et al (2015) Shape-dependent skin penetration of silver nanoparticles: does it really matter? Sci Rep 5:16908. https​://doi.org/10.1038/srep1​6908 52. Turkevich J, Kim G (1970) Palladium: preparation and catalytic properties of particles of uniform size. Science 169:873–879. https​://doi.org/10.1126/scien​ce.169.3948.873 53. Kimling J, Maier M, Okenve B et  al (2006) Turkevich method for gold nanoparticle synthesis revisited. J Phys Chem B 110:15700–15707. https​://doi.org/10.1021/jp061​667w 54. Labouta HI, Liu DC, Lin LL et al (2011) Gold nanoparticle penetration and reduced metabolism in human skin by toluene. Pharm Res 28:2931–2944. https​://doi.org/10.1007/s1109​5-011-0561-z 55. Labouta HI, El-Khordagui LK, Kraus T, Schneider M (2011) Mechanism and determinants of nanoparticle penetration through human skin. Nanoscale 3:4989. https​://doi.org/10.1039/c1nr1​ 1109d​ 5 6. Huang Y, Yu F, Park Y-S et al (2010) Co-administration of protein drugs with gold nanoparticles to enable percutaneous delivery. Biomaterials 31:9086–9091. https​://doi.org/10.1016/j.bioma​ teria​ls.2010.08.046 57. Hong S, Bielinska AU, Mecke A et al (2004) Interaction of poly(amidoamine) dendrimers with supported lipid bilayers and cells: hole formation and the relation to transport. Bioconjug Chem 15:774–782. https​://doi.org/10.1021/bc049​962b 58. Kirjavainen M, Urtti A, Jääskeläinen I et  al (1996) Interaction of liposomes with human skin in  vitro—the influence of lipid composition and structure. Biochim Biophys Acta Lipids Lipid Metab 1304:179–189. https​://doi.org/10.1016/S0005​-2760(96)00126​-9 59. Shanmugam S, Song C-K, Nagayya-Sriraman S et  al (2009) Physicochemical characterization and skin permeation of liposome formulations containing clindamycin phosphate. Arch Pharm Res 32:1067–1075. https​://doi.org/10.1007/s1227​2-009-1713-0 60. Lee O, Jeong SH, Shin WU et  al (2013) Influence of surface charge of gold nanorods on skin penetration. Ski Res Technol 19:e390–e396. https​://doi.org/10.1111/j.1600-0846.2012.00656​.x 61. Mahmoud NN, Al-Qaoud KM, Al-Bakri AG et  al (2016) Colloidal stability of gold nanorod solution upon exposure to excised human skin: effect of surface chemistry and protein adsorption. Int J Biochem Cell Biol 75:223–231. https​://doi.org/10.1016/j.bioce​l.2016.02.020 62. Hao F, Jin X, Liu QS et al (2017) Epidermal penetration of gold nanoparticles and its underlying mechanism based on human reconstructed 3D episkin model. ACS Appl Mater Interfaces 9:42577–42588. https​://doi.org/10.1021/acsam​i.7b137​00 63. Suk JS, Xu Q, Kim N et  al (2016) PEGylation as a strategy for improving nanoparticle based drug and gene delivery. Adv Drug Deliv Rev 99:28–51. https​://doi.org/10.1016/j. addr.2015.09.012 64. Tsai H-C, Hsiao PF, Peng S et al (2016) Enhancing the in vivo transdermal delivery of gold nanoparticles using poly(ethylene glycol) and its oleylamine conjugate. Int J Nanomed 11:1867–1878. https​://doi.org/10.2147/IJN.S1025​99 65. Mahmoud NN, Alkilany AM, Dietrich D et al (2017) Preferential accumulation of gold nanorods into human skin hair follicles: effect of nanoparticle surface chemistry. J Colloid Interface Sci 503:95–102. https​://doi.org/10.1016/j.jcis.2017.05.011 66. Labala S, Mandapalli PK, Kurumaddali A, Venuganti VVK (2015) Layer-by-layer polymer coated gold nanoparticles for topical delivery of imatinib mesylate to treat melanoma. Mol Pharm 12:878– 888. https​://doi.org/10.1021/mp500​7163 67. Bessar H, Venditti I, Benassi L et al (2016) Functionalized gold nanoparticles for topical delivery of methotrexate for the possible treatment of psoriasis. Colloids Surf B Biointerfaces 141:141–147. https​://doi.org/10.1016/j.colsu​rfb.2016.01.021 68. Pan A, Zhong M, Wu H et al (2018) Topical application of keratinocyte growth factor conjugated gold nanoparticles accelerate wound healing. Nanomed Nanotechnol Biol Med 14:1619–1628. https​ ://doi.org/10.1016/j.nano.2018.04.007 69. Niu J, Chu Y, Huang Y-F et al (2017) Transdermal gene delivery by functional peptide-conjugated cationic gold nanoparticle reverses the progression and metastasis of cutaneous melanoma. ACS Appl Mater Interfaces 9:9388–9401. https​://doi.org/10.1021/acsam​i.6b163​78

13

234

Reprinted from the journal

Topics in Current Chemistry (2020) 378:8



70. Chen Y, Wu Y, Gao J et  al (2017) Transdermal vascular endothelial growth factor delivery with surface engineered gold nanoparticles. ACS Appl Mater Interfaces 9:5173–5180. https​://doi. org/10.1021/acsam​i.6b159​14 71. Safwat MA, Soliman GM, Sayed D, Attia MA (2017) Gold nanoparticles capped with benzalkonium chloride and poly (ethylene imine) for enhanced loading and skin permeability of 5-fluorouracil. Drug Dev Ind Pharm 43:1780–1791. https​://doi.org/10.1080/03639​045.2017.13390​82 72. Safwat MA, Soliman GM, Sayed D, Attia MA (2018) Fluorouracil-loaded gold nanoparticles for the treatment of skin cancer: development, in vitro characterization, and in vivo evaluation in a mouse skin cancer xenograft model. Mol Pharm 15:2194–2205. https​://doi.org/10.1021/acs.molph​armac​ eut.8b000​47 73. Boca S, Berce C, Jurj A et al (2017) Ruxolitinib-conjugated gold nanoparticles for topical administration: an alternative for treating alopecia? Med Hypotheses 109:42–45. https​://doi.org/10.1016/j. mehy.2017.09.023 74. Anirudhan TS, Nair SS, Sasidharan AV (2017) Methacrylate-stitched β-cyclodextrin embedded with nanogold/nanotitania: a skin adhesive device for enhanced transdermal drug delivery. ACS Appl Mater Interfaces 9:44377–44391. https​://doi.org/10.1021/acsam​i.7b166​86 75. Anirudhan TS, Nair SS, Sekhar VC (2017) Deposition of gold-cellulose hybrid nanofiller on a polyelectrolyte membrane constructed using guar gum and poly(vinyl alcohol) for transdermal drug delivery. J Memb Sci 539:344–357. https​://doi.org/10.1016/j.memsc​i.2017.05.054 76. Crisan D, Scharffetter-Kochanek K, Crisan M et  al (2018) Topical silver and gold nanoparticles complexed with Cornus mas suppress inflammation in human psoriasis plaques by inhibiting NF-κB activity. Exp Dermatol 27:1166–1169. https​://doi.org/10.1111/exd.13707​ 77. Wang S, Yan C, Zhang X et al (2018) Antimicrobial peptide modification enhances the gene delivery and bactericidal efficiency of gold nanoparticles for accelerating diabetic wound healing. Biomater Sci 6:2757–2772. https​://doi.org/10.1039/C8BM0​0807H​ 78. Fratoddi I, Benassi L, Botti E et al (2019) Effects of topical methotrexate loaded gold nanoparticle in cutaneous inflammatory mouse model. Nanomed Nanotechnol Biol Med 17:276–286. https​://doi. org/10.1016/j.nano.2019.01.006 79. Hernández Martínez SP, Rivera González TI, Franco Molina MA et al (2019) A novel gold calreticulin nanocomposite based on chitosan for wound healing in a diabetic mice model. Nanomaterials 9:75. https​://doi.org/10.3390/nano9​01007​5 80. Mandal B, Rameshbabu AP, Soni SR et al (2017) In situ silver nanowire deposited cross-linked carboxymethyl cellulose: a potential transdermal anticancer drug carrier. ACS Appl Mater Interfaces 9:36583–36595. https​://doi.org/10.1021/acsam​i.7b107​16

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Reprinted from the journal

235

13

Topics in Current Chemistry (2020) 378:43 https://doi.org/10.1007/s41061-020-00306-6 REVIEW

Proteins‑Based Nanocatalysts for Energy Conversion Reactions Daily Rodriguez‑Padron1 · Md Ariful Ahsan2,3 · Mohamed Fathi Sanad2 · Rafael Luque1,4 · Alain R. Puente Santiago2  Received: 2 January 2020 / Accepted: 10 June 2020 © Springer Nature Switzerland AG 2020

Abstract In recent years, the incorporation of molecular enzymes into nanostructured frameworks to create efficient energy conversion biomaterials has gained increasing interest as a promising strategy owing to both the dynamic behavior of proteins for their electrocatalytic function and the unique properties of the synergistic interactions between proteins and nanosized materials. Herein, we review the impact of proteins on energy conversion fields and the contribution of proteins to the improved activity of the resulting nanocomposites. We address different strategies to fabricate proteinbased nanocatalysts as well as current knowledge on the structure–function relationships of enzymes during the catalytic processes. Additionally, a comprehensive review of state-of-the-art bioelectrocatalytic materials for water-splitting reactions such as hydrogen evolution reaction (HER) and oxygen evolution reactions (OER) is afforded. Finally, we briefly envision opportunities to develop a new generation of electrocatalysts towards the electrochemical reduction of ­N2 to ­NH3 using theoretical tools to built nature-inspired nitrogen reduction reaction catalysts. Keywords  Proteins · Nanomaterials · Energy conversion Daily Rodriguez-Padron and Md Ariful Ahsan contributed equally to this work.

Chapter 8 was originally published as Rodriguez‑Padron, D., Ahsan, M. A., Sanad, M. F., Luque, R. & Santiago, A. R. P. Topics in Current Chemistry (2020) 378: 43. https://doi.org/10.1007/s41061-02000306-6. * Alain R. Puente Santiago [email protected] 1

Departamento de Química Orgánica, Universidad de Córdoba, Campus de Rabanales, Edificio Marie Curie (C‑3), Ctra Nnal IV‑A, Km 396, 14014 Córdoba, Spain

2

Department of Chemistry, University of Texas at El Paso, 500 W. University Avenue, El Paso, TX 79968, USA

3

Nanosystems Engineering Research Center for Nanotechnology-Enabled Water Treatment, Houston 77005, USA

4

Peoples Friendship University of Russia, (RUDN University), 6 Miklukho‑Maklaya Str., 117198 Moscow, Russia



Reprinted from the journal

237

13

Vol.:(0123456789)

Topics in Current Chemistry (2020) 378:43

1 Introduction Proteins are well-designed nanosized machines that can dynamically catalyze myriad catalytic reactions with molecular-level accuracy [1–4]. As an example of heterogeneous catalysis, electrocatalysis, which explores the relationship between the physicochemical properties of electrode materials and both their underlying mechanism and rate of the electrode reactions, have been widely investigated. Particularly, the development of bio-based advanced energy conversion nanosystems has recently emerged as an important and attractive topic. Water-splitting reactions represent a promising and sustainable way to obtain hydrogen and oxygen through the development of renewable energy fuel devices, which fulfill a crucial role in solving the global energy crisis. The mechanistic principles that proteins use to acts as efficient catalysts towards the generation of fuels have been intensively investigated [5]. As a result, bio-inspired synthetic catalysts have been generated to deeply study the biological activity of different kinds of enzymes for the production of added-value chemical compounds. Additional efforts have been performed to unravel the connection between the structure and the catalytic performances of active enzymes for water-splitting reactions [6–9]. It has been established that there is a good relationship between the structural or conformational changes of enzymes and their catalytic activity on the ms–μs timescale. [10–15] In this sense, small conformational variations or motions of just a few residues in the three-dimensional (3D) protein frameworks can also improve the efficiency of the electrochemical reaction that occurs at the electrode–electrolyte interface and, in turn, the overall electrocatalytic process. The impressive abilities of proteins to adopt catalytically competent configurations via efficient conformational changes on specific domains of their tertiary structures towards environmental changes such as pH, temperature or even ionic strength then become promising candidates for the development of high-performance electrocatalysts. Although several enzymes exhibit high catalytic performances, electronic conductivity through the amino acids of the tertiary structure is still being low, which represents a drawback to obtaining effective electrocatalytic nanosystems [16–19]. To address this limitation, several strategies have been attempted in the past years. Among them, the engineering of nanobiointerfaces at the molecular level using different immobilization approaches to optimize the efficiency of electrocatalytic reactions has become one of the most effective strategies [20]. To develop the latter strategy, a couple of key factors have been fairly well determined. Firstly, the enzyme immobilization process should be optimized to reach a high coverage of active molecules on the electrode surface. Secondly, the electronic wiring between the active sites of the immobilized enzyme molecules and the electrodes plays a crucial role in generating advanced electrocatalysts. Consequently, they have been integrated via covalent or non-covalent functionalization to a large number of conductive nanomaterials (i.e., gold nanoparticles, graphene sheets, and nanotubes) to generate the most electroactive orientations of active centers assuring faster electron transfer (ET) processes and enhancing the conductive wiring and, in turn, the catalytic efficiency of the resulting biomaterials [21–26].

13

238

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





N-doped carbon nanostructures have been widely employed as high-performance electrocatalysts for water-splitting reactions. Particularly, the development of N‐doped graphene and N‐doped carbon nanotubes has sparked a lot of interest in recent years in the material science community. In this direction, Junji Nakamura and coworkers, through a seminal work, have elucidated, after a long period of controversy, the mechanism of the oxygen reduction reaction (ORR) in N-doped carbon materials. They discovered that pyridinic N can generate carbon atoms with Lewis basicity, which leads to the adsorption of ­O2 molecules onto them at the first step of oxygen electroreduction [27, 28]. Finding new avenues to synthesize metal-free electrocatalysts, proteins have been used as a N-source to create very promising N-doped carbon nanocatalysts due to the significant contents of pyrrolic N, pyridinic N and amine groups in their molecular architectures. Among the most significant contributions found is the covalent attachment of hemoglobin to fructose-functionalized graphene oxide nanostructures to built high-performance ORR electrocatalysts [6]. Through an elegant and pioneering work, Santiago and coworkers have developed a methodology to multiply the electrocatalytic activity of graphene-oxide nanoplatforms through the unfolding of adsorbed hemoglobin molecules using fructose linkers as denaturing agents. During the denaturalization process, the proteins adopt a fibrin-like structure, releasing almost all their redox-active centers and exposing their pyridine groups to electrochemical interfaces. The fixed hemoglobins significantly boosted the ORR properties of the graphene oxide nanoplatforms in terms of onset potential and current density. This work paved the way towards the development of metalfree nanobioelectrocatalysts. Inspired by nature, Compton and coworkers developed a revolutionary methodology to amplify the ORR bioelectrocatalytic signal of redox proteins using layered structured films composed of conductive polymers and hemoglobin molecules in a sandwich-like configuration [29]. The synergistic interactions between Nafion and hemoglobin gave rise to the full conversion of oxygen to water by a four-electron pathway of the water-splitting process, which, in turn, boosted electrocatalytic activity. Redox enzymes possess myriad different types of redox sites, including hemes, chlorins, quinones, favins, Fe–S clusters, tyrosine and tryptophan residues, copper, molybdenum, and manganese ions, which can catalyze a large portfolio of redox reactions based on their electron-transporting properties and structural affinity for the reactant molecules [30–34]. Remarkably, they could be easily integrated into different types of nanoplatforms to significantly boost their electrocatalytic performances. In this review, we provide a timely summary of recent methodologies developed to fabricated high-performance protein-based nanoelectrocatalysts for water-splitting reactions. We describe the relationship between the structural properties of the nanobiomaterials and their electrocatalytic performances based on an understanding of the reaction mechanisms and the role of the biomolecule structure. Additionally, we envision the development of molecular biocatalytic nitrogen reduction reaction (NRR) nanomaterials for the efficient electrochemical production under mild conditions of ammonia, which is used as a refrigerant gas, for purification of water Reprinted from the journal

239

13





Topics in Current Chemistry (2020) 378:43

supplies, and in the manufacture of plastics, explosives, textiles, pesticides, dyes, and other chemicals.

2 Proteins as Efficient Building Blocks to Construct Advanced Electrocatalytic Materials Currently, many researchers, motivated by the high efficiency observed in nature, have directed their studies towards the design of bioinspired materials. In this sense, enzymes are highly efficient and selective biological catalysts. The presence of redox groups in this type of biomolecule has led to different electrochemical applications. Enzymes can catalyze a series of redox reactions with great technological importance, which generally require high overpotentials [35]. It is worth highlighting that the use of enzymes provides many advantages, such as high electrocatalytic currents and the possibility of carrying out electrocatalytic processes in a reversible way. For instance, hydrogenases are metalloenzymes, which catalyze the reversible conversion of hydrogen [36]. Depending on the metal content in the active sites of the protein, three main classes of hydrogenases can be highlighted: [FeFe], which contains two iron atoms and is a faster biological catalyst for hydrogen oxidation/ reduction reactions; [NiFe], which possesses an active heterobimetallic site; and [Fe], which contains only one iron atom in its structure. Due to their ability to efficiently electrocatalyze hydrogen production, a series of hydrogen fuel cells has been designed using these enzymes, representing a valuable alternative in the design of biodegradable energy devices [37–39]. Hemoglobin, in particular, is a protein that contains iron atoms in its structure, and that has been widely characterized due to its importance for living beings [40]. Hemoglobin comprises a redox, globular, and tetrameric structure. Each subunit is formed by polypeptide chains, mostly in alpha-helix conformation, that encompass in their structure a heme group containing an atom of Fe. This ­Fe2+ ion is in the center of an organic heterocycle called porphyrin, which is responsible for reacting with oxygen and carrying out blood oxygen transport. These characteristics allow these types of proteins to act as smart catalytic platforms in the oxygen evolution reaction, as well as in other types of oxidation reactions. Moreover, peroxidases are another class of enzymes that have been used as sustainable electrocatalysts. These ­ 2O2, thanks to the presproteins catalyze the oxidation of organic molecules using H ence of the Fe-porphyrin group [41]. In particular, enzymes known as multicopper oxidases (MCOs) show excellent properties towards electroreduction of O ­ 2 at high potentials [42]. MCOs are a family of metalloenzymes that possess three different copper sites: Cu types 1, 2 and 3. In this case, oxidation of the reducing substrate occurs at the Cu type 1 site (T1), while the reduction of ­O2 occurs in the trinuclear group T2/T3 [43]. A wide range of strategies has been used to immobilize these enzymes, from covalent bonding to carbon nanotubes and graphene conductive nanomaterials to encapsulation in silica structures [44, 45]. Furthermore, it is known that specific enzymes exhibit excellent ability to reduce ­CO2–CO and formic acid, behaving as reversible electrocatalysts. For example, the

13

240

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43



carbon monoxide dehydrogenase enzymes present in anaerobic organisms contain a group (NiFe-4 S) that catalyzes the interconversion of C ­ O2 and CO [41]. The enzyme formate dehydrogenase, with active sites of pyranopterin molybdenum/ tungsten, can catalyze the interconversion of C ­ O2 and formate [45]. In particular, formate has attracted much interest as an energy source, since it is easier to store and transport than hydrogen. Using a bio-synthetic approach, many scientists have developed models to mimic metalloenzymes such as hydrogenases and nitric oxide reductase, among others [46–48]. Although the catalytic efficiencies obtained for these artificial systems are still low in comparison with natural systems, such strategies are very promising and have great potential for the future.

3 Influence of Protein Conformational Variations on the Electrocatalytic Activity The structure–function relationship of protein-based nanocatalysts has been investigated thoroughly in order to gain insights into the rules that govern the bioelectrocatalytic process at the molecular scale. The understanding of bioelectrocatalytic mechanisms at a structural level constitutes vital knowledge towards the design of enzyme-based nanomaterials as potential electrocatalysts for bioenergy applications. Nevertheless, although the catalytic activity of various redox proteins has been studied widely and linked with their structural properties, the influence of structural modifications to the protein architecture on the bioelectrocatalytic properties of protein-based nanosystems have been scarcely reported. In this direction, Alina Sekretaryova and colleagues have studied the oxygen electroreduction behavior of single redox enzymes when they impact with ultramicroelectrode surfaces (Fig.  1) [49]. Interestingly, they found that the electrochemical signals followed spike-shaped patterns instead of steady-current steps, which could most likely be associated with the partial denaturation or structural changes of single laccase molecules during the adsorption process. In this regard, the conformational Fig. 1  Scheme showing the principle of detection of the catalytic current from a single enzyme molecule

Reprinted from the journal

241

13





Topics in Current Chemistry (2020) 378:43

or structural variations of the proteins might be promoting direct electron transfer (DET) between the electrode and the T1 redox centers of the enzymes, and, therefore, the electroreduction of molecular oxygen to water. Despite the aforementioned breakthrough, there is a lack of knowledge of the structural and conformational properties of enzymes at each catalytic step. Therefore, the chemical and structural factors that enable electrochemical activity, as well as a deep understanding of the electron transfer processes associated with the catalytic reaction, has yet to be completely unraveled experimentally. In this direction, amplifying the low electrocatalytic signal of enzyme-electrode signals using lowdimensional materials to improve the electronic communication of proteins with the surface electrodes and, in turn, facilitate the electrochemical investigation of the catalytic role of enzymes, represents a suitable alternative. Also, a large number of advanced electrochemical techniques, quartz-crystal microbalance measurements, and spectroscopic techniques such as resonance Raman and infrared spectroscopy should be coupled in situ to the resulting nanocomposites to gain insightful advances in the mechanistic understanding of enzyme-based electrocatalysts [2]. As a first approach, Rafael Luque and coworkers have reported that the partially unfolded states of laccase molecules mechanochemically immobilized on magnetic nanoparticles give rise to very low onset potential and outstanding currents towards the direct bioelectrocatalytic reduction of molecular oxygen [50]. The low content of α-helix, together with the increase in the number of low-frequency β-sheets, was associated directly with improved electrocatalytic activity (Fig.  2). This pioneering work sheds light on the nature of the nanometric interaction between proteins and low-dimensional materials and the impact on their bioelectrocatalytic function. However, the detailed mechanism of how conformational variations of proteins can lead to improvements in electrocatalytic activity is still unclear. Therefore, investigation of the underlying chemistry of protein-based nanocatalysts in the electrocatalytic process is still in its infancy. A lot of research should be carried out to disentangle the structural variations that proteins attached to lowdimensional materials undergo under electrocatalytic conditions. Molecular dynamics studies of the movement of these proteins during bioelectrocatalysis can be applied successfully to unravel their dynamic catalytic performances at the molecular scale. Additionally, control of the structure–function properties of nanobiocatalyts using molecular biology approaches such as the replacement of specific amino acids to greatly improve their electrocatalytic yields constitutes a promising strategy. Briefly, we envision that this knowledge will open the door towards the development of a new generation of biomaterials with unbeatable electrocatalytic properties.

4 Protein Based‑Materials as High‑performance Water‑Splitting Electrocatalysts Proteins have ultimately inspired the design of powerful nanomaterials towards the electrochemical reduction of ­O2. Two approaches have delivered impressive outcomes: (1) N-based nanocomposites fabricated using N-containing proteins and (2) the incorporation of redox-active sites into the nanohybrid structures.

13

242

Reprinted from the journal

Fig. 2  Fourier-transform infrared (FT-IR) spectroscopy of a laccase functionalized waste-derived iron oxide nanoparticles (LAC-DA-Fe2O3) and b deconvoluted FT-IR spectrum of LAC-DA-Fe2O3 nanobioconjugates (reprinted from ref. [50])

Topics in Current Chemistry (2020) 378:43

Reprinted from the journal



243

13





Topics in Current Chemistry (2020) 378:43

4.1 N‑Containing Proteins ORR-based electrocatalysts are considered as essential components of metal–air batteries and fuel cells [51, 52]. The most common ORR electrocatalysts are mainly platinum and its alloys due to their higher current density and relatively lower overpotential. But the higher cost, availability, anode crossover and severe intermediate tolerance, etc. of platinum impose restrictions on the advancement of fuel cell technologies. Therefore, the development of low-cost, as well as high-performance ORR catalysts, has being investigated in recent times [53]. In this regard, low-cost and metal-free carbon-based electrocatalysts have drawn immense attention owing to their superior durability and excellent electrocatalytic activity. There are different types of carbon materials, including porous carbon, graphene, activated carbon and carbon nanotubes, which have been used to improve the catalytic activity and durability of electrocatalysts by maximizing their electroactive surface area. Recently, it was found that the electrocatalytic performance of the ORRs can be enhanced by nitrogen doping of carbon materials [28]. The presence of abundant and free-flowing sp2 hybridized pi electrons are mainly responsible for this significant enhancement in performance. Lone pair electrons of nitrogen atoms also help to connect the delocalized conjugated systems, thereby resulting in additional improvement in electrocatalytic activity. Nitrogen doping of carbon nanostructures can be accomplished by two conventional procedures such as ‘in situ doping,’ which is done during the synthesis of carbon materials, and ‘post doping,’ which is achieved through the post-treatment of carbon nanostructures. But these methods usually require expensive hardware, multi-step processes and high energy consumption, which eventually restricts their practical applicability. To overcome these problems, exploitation of N-containing proteins can be beneficial because of their easy availability as well as the simplicity of preparing N-doped carbonaceous materials. Blood proteins (BPs) from animals, which can be sourced from the meat industry, contain various amino acids such as proline, glutamic acid and tyrosine, and an abundance of hemoprotein. These substances can successfully produce active electrocatalysts for ORR in both acidic and alkaline environments. Therefore, BP could be used as a potential precursor material for making highly active electrocatalysts. Guo et  al. [54] reported a novel approach for designing nitrogen-enriched carbonbased electrocatalysts for ORR through the co-pyrolysis of a carbon black support and BP. At first, the BP was decomposed at 350 °C for 5 h under a constant flow of nitrogen. The pyropolymer obtained was then mixed with the carbon black support by ball milling. The sample yielded was again treated at 1000 °C for 2 h in a nitrogen atmosphere (BP350C1000) and then used as an electrocatalyst for ORR. As a control, the BP was also carbonized for 2  h at 1000  °C and the sample was denoted as BP1000. Furthermore, BP3501000 was prepared by the continuous heat treatment of BP for 5  h at 350  °C and 2  h at 1000  °C without adding the carbon black into it. The results demonstrated that the nitrogen present in the as-synthesized electrocatalyst from BP is mainly in the form of pyrrolic- and pyridinic-type nitrogen species. The electrocatalyst comprising higher amounts of pyrrolic-type nitrogen species exhibited better electrocatalytic performance towards the ORR in terms of limited current density, half-wave potential, and onset potential. In this

13

244

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43



study, the best electrocatalyst (BP350C1000) exhibited an onset potential of 0.90 V and a half-wave potential of 0.78 V, which are close to the Pt/C, and catalyzes the electrochemical reduction of ­O2 through a four-electron pathway (Fig. 3a–d). It was also found that the carbon black support plays a vital role in the pyrolysis process, resulting in improved catalytic activity. The difference between BP350C1000 and BP3501000 in the electrocatalytic activity of the ORR can be ascribed to three features: (1) introducing carbon support yields more exposed electrocatalytic active sites and delivers a greater space for more catalytically active sites by preventing aggregation during the pyrolysis process; (2) addition of carbon black support helps to create new catalytic active sites on modified carbon materials with other formed N-containing groups from the further decomposition of BP350 precursors during the high-temperature carbonization process, and (3) a larger amount of pyrrolic-N configuration is formed in BP350C1000 than in BP3501000. These results indicated that the utilization of carbon black as a conductive material and inserting a matrix to create catalytic active sites in the pyrolysis process can significantly enhance its ORR electrocatalytic activity in alkaline media. Collagen—the most abundant extracellular protein—is usually found in mammalian connective tissues. Animal skin wastes are used as a potential precursor to extract collagen easily, and the collagen obtained can also be used as an efficient precursor for the preparation of carbon nanostructures. Ajayan et al. [55] reported a

Fig. 3  a Oxygen reduction reaction (ORR) polarization curves for blood proteins (BP) BP1000, BP3501000, and BP350C1000 before and after accelerated aging tests (AAT) at a rotation rate of 1600  rpm. b Tafel plots for BP1000, BP3501000, and BP350C1000. c ORR polarization curves for BP350C1000 at different rotation rates. d Koutecky-Levich curves (reprinted from Ref. [54]) Reprinted from the journal

245

13





Topics in Current Chemistry (2020) 378:43

facile method to prepare nitrogen-rich carbon nano-onions from collagen to use as an ORR electrocatalyst. They used goat skin wastes as a source of collagen. Electrochemical measurements demonstrated that the electrocatalytic activity of the prepared carbon samples is significantly comparable to that of commercially available 20% Pt/C electrocatalyst. The best electrocatalytic performance was observed for the carbon sample synthesized at 750 °C for 8 h with an onset potential of  −50 mV vs. Ag/AgCl while comparing to 81 mV for 20% Pt/C. The electrocatalyst with higher pyridinic nitrogen content exhibited an effective four-electron transfer ORR mechanism with a current density that is almost comparable to commercial Pt/C. The main driving force enhancing ORR performance is the existence of the pyridinic nitrogen atoms, which act as active sites as well as delivering net positive charge on the adjacent carbon atoms to facilitate oxygen adsorption along with the attraction of electrons from the anode. Furthermore, the nitrogen-rich carbon nano-onion-based electrocatalysts displayed outstanding durability performance in the alkaline medium as well as superior methanol tolerance as compared with the 20% Pt/C. Laccases are known as MCOs that are usually found in bacteria, fungi, and plants. They can couple the one-electron oxidation of four substrate equals with four-electron reduction of an oxygen molecule to water. The electroreduction mechanism happens by following a ping-pong mechanism. Oxidation of the substrates takes place near the solvent-accessible T1 site. The electrons are then transferred through the protein by following the Cys–His pathway over a distance of ∼12 Å to the trinuclear copper center where the ORR takes place. Luque et  al. [56] reported the synthesis of an unprecedented electrically active silica-encapsulated laccase material through a facile and eco-friendly one-pot biosilicification process and used for the ORR for the first time. The one-pot biosilicification preparation facilitated the lodging of the enzymes in the highly active alignments required for direct transfer of the electrons of T1 redox centers. As a result, the biosilicified laccase material that is deposited on the nickel electrodes displayed an effective bioelectrocatalytic reduction of oxygen, delivering an outstanding current density of up to 0.94 mA/cm2 and good long-term stability properties (Fig. 4). To design a metal-free ORR electrocatalyst, Chen et  al. [57] reported a novel strategy for preparing nitrogen-doped carbon nanomaterial-based electrocatalysts by utilizing pyrolysis of the protein-enriched enoki mushroom at 900  °C, along with carbon nanotubes as an inserting matrix and conductive agent. The mushrooms are an abundantly available, renewable biomass source and are amino acid-rich. They behave as a single precursor material for both carbon and heteroatoms, thus avoiding the utilization of complicated chemicals in the preparation method. It was found that numerous forms of nitrogen (graphitic, pyrrolic and nitrile) were inserted into the carbon molecular skeleton of the product, which displayed excellent ORR electrocatalytic performance as well as better durability behavior in alkaline medium compared with those in acidic medium. The onset potential of the electrocatalyst obtained in this study was about 0.94 V in the alkaline medium, which is comparable with state-of-the-art Pt/C electrocatalysts (0.98 V). Remarkably, the measured ORR half-wave potential of the prepared electrocatalyst was found to be 0.81 V in alkaline medium, which is slightly lower than that of the 20 wt% Pt/C electrocatalysts

13

246

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





Fig. 4  Cyclic voltammograms of a free laccase and b biosilicified laccase under oxygen purging in 0.1 M phosphate-buffered saline (PBS) at pH 6. Scan rate: 0.01 V/s. b Chronoamperometric responses at Eapp = 0.2 V (reprinted from Ref. [56])

(0.86  V). Also, the reduction reaction followed the four-electron transfer ORR mechanism including the direct reduction pathway. The authors suggested that the graphitic-nitrogen species that are mostly responsible for the ORR performance can also function as an active center for the ORR reaction, whereas the pyrrolic-nitrogen species can behave as an active promoter for ORR only. Nature offers many biological materials such as proteins, enzymes, etc. that contain histidines, which could be utilized as a natural source of “pyridine-like” nitrogen species offering catalytically active sites for ORR. Among such proteins, hemoglobin (Hb) is a tetrameric redox protein comprised of four polypeptide chains wherein each polypeptide chain encircles at least one active iron redox center. Puente-Santiago et al. [6] reported an unprecedented bottom-up method to prepare a high-performance bioelectrocatalytic system for ORR from the covalent anchorage of Hb to fructose-modified graphene oxide nanoplatforms via the glycosylation reaction mechanism. The resulting nanobiomaterials were prepared at room temperature (GO-Fruc@Hb-RT) as well as at 80 °C (GO-Fruc@Hb-HT) to gain information on the structural characteristics of the redox enzymes in the prepared nanomaterials as well as their electrocatalytic activity toward ORR reactions. Electrocatalytic experiments revealed that the measured onset potentials for both three-component nanocomposites moved to more positive values than their individual parts, demonstrating their higher electrocatalytic activities. Furthermore, the limiting current density was also increased for both nanocomposites, with the best electrocatalytic activity for the nanocomposite being synthesized at a higher temperature because the number of pyridinic nitrogen species is increased at higher temperature due to the additional conformational changes. Additionally, a few examples in the literature have addressed the fabrication of protein-based nanosystems as high-performance HER catalysts. One outstanding strategy is the synthesis of engineered biocatalysts for ­H2 evolution. In this sense, Bren and coworkers [58] have elegantly synthesized a synthetic protein named cobalt mimochrome VI*, which is able to remarkably catalyze the hydrogen Reprinted from the journal

247

13





Topics in Current Chemistry (2020) 378:43

evolution process, delivering an ultralow onset potential and a turnover number exceeding 230,000. Also, Santiago and coworkers have developed an innovative HER-based catalyst composed of graphitic carbon nitride nanostructures functionalized with different silver contents and α-rich proteins. The as-synthesized nanomaterials rendered an impressive overpotential of 79 mV at a current density of 10 mA/ cm2 towards molecular hydrogen production, which is comparable with the most efficient HER electrocatalysts reported in the literature [59]. 4.2 Redox Proteins Current energy requirements have demanded that we find new alternatives for the development of catalytic systems, which be able to carry out proton-coupled redox reactions. In particular, ORR has been a focus of attention of the scientific community in recent years, since it is the cathodic reaction of fuel cells [60, 61]. Indeed, ORR is crucial to the function of hydrogen fuel cells. Therefore, one of the biggest current challenges is the preparation of efficient catalysts for ORRs, particularly for regenerative or reversible fuel cells, which are fuel cells, and, in a reversible way, can produce ­H2 and ­O2, by water electrolysis [62]. Despite the enormous progress accomplished so far, the development of electrocatalysts with high activity and low costs is still a great challenge. The current bottleneck of fuel cells is ORR reactions, which are the limiting step for the generation of electricity. So far, the most used systems for ORR reactions are mainly platinumbased materials (or their alloys). However, due to the high cost of Pt, alternative catalysts based on other, less expensive, metals, such as transition metals, as well as other non-metal-containing materials are being actively sought [63]. The last few decades have shown considerable improvements in the design of efficient catalysts for ORR. An important factor that should be considered when choosing a good catalytic system is selectivity towards the 4­ H+/4e− reduction of O ­ 2 to ­H2O, as opposed to the ­2H+/2e− reduction of ­O2 to ­H2O2. Besides, the reduction of ­O2 by one electron, to yield ­O2−, is also undesired [64]. In this regard, in nature, biological systems possess specific proteins that favor the selective ­4H+/4e− reduction of O ­ 2. For instance, cytochrome c oxidase ©cO) catalyzes ­O2 reduction as part of the respiratory complex that drives adenosine triphosphate (ATP) biosynthesis. The active sites of CcO enzymes involve a Fe-containing heme (called heme α3), a distal Cu, and a post-translationally modified tyrosine amino acid (tyrosine 244) [65]. Another example of O ­ 2-reducing metalloproteins is the family of MCOs. In this case, the mechanism of the oxygen reduction reactions in MCOs favors the direct ­4e−/4H+ reduction process of ­O2 to ­H2O. It is worth mentioning that some MCOs mediate ORR close to the thermodynamic potential of the ­O2/H2O couple [66–69]. Several materials have been designed employing proteins that can act as active catalysts for ORR. For example, synthetic models of iron-porphyrins, incorporating a distal CuI center connected via a meso-position with a pendant ligand that resembles the active site coordination in CcO have been described. Such samples have been employed in ORR by immobilization on graphite disk electrodes. As well,

13

248

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





mononuclear iron porphyrins adsorbed onto graphite surfaces have been employed successfully as catalysts for the mentioned reaction through a 4­ H+/4e− pathway. Immobilization of iron porphyrin catalysts onto edge plane graphite shifts the potential of FeIII/II redox couple to more positive values than those observed in homogeneous solutions, and, therefore, results in lower ORR overpotentials, in comparison to the homogeneous cases [70–73]. Furthermore, cobalt-porphyrin based catalysts have also been widely recognized as efficient electrocatalysts for oxygen electroreduction reactions. Nonetheless, such systems possess disadvantages related to the poor selectivity towards H ­ 2O, obtaining predominantly ­H2O2. In turn, Co-porphine immobilized onto edge plane graphite surface have shown ­4e− reduction of ­O2 under air-saturated 1  M ­HClO4 solution [74]. A recombinant enzyme, namely CotA laccase has been also employed successfully as a bioelectroctalytic system for ORR (Fig.  5). The metalloenzymes were immobilized onto citrate-coated gold nanoparticles (AuNPs) through attractive electrostatic interactions. Interestingly, electronic wiring of the enzymes via a T2/ T3 trinuclear with the T2/T3 redox groups facing the surfaces of the electrodes was observed. This work combined, for the first time, both advances in the synthesis of recombinant redox enzymes and the advantages of colloidal nanosystem synthesis to create electrostatically self-assembled nanomaterials with remarkable ORR properties [75].

5 Proteins‑Inspired Advanced NRR Electrocatalysts As is widely known, ammonia is very important to the global economy as a fertilizer feedstock and household chemical as well as a desirable refrigerant in industry; it is also a chemical precursor in addition to being considered a future fuel alternative [76]. Currently, the incumbent Haber–Bosch process (HBP) is the strategy most widely used to synthesize ammonia for industrial-scale production. Due to its large number of limitations, including high cost, energy consumption and process complexity, finding other alternatives is needed [77]; therefore, progress in the fields of biocatalysis and electrocatalysis to understand the electrochemical reduction of dinitrogen ­(N2) to ammonia (­ NH3) and enable a greener path to ammonia production

Fig. 5  Schematic diagrams of a spore coat protein A (CotA) laccase immobilized onto citrate-coated gold nanoparticles (AuNPs) and b ORR processes onto the nanobiomaterial surfaces (reprinted from Ref. [75]) Reprinted from the journal

249

13





Topics in Current Chemistry (2020) 378:43

need to be developed. Electrochemical reduction of nitrogen to ammonia is thermodynamically predicted to be more efficient than the HBP by about 20%. Besides, an electrochemical process could provide the advantage of eliminating fossil fuels as the precursor of ­H2 and energy by the use of water molecules for hydrogen production [78]. In this manner, electrochemical systems offer additional benefits, including scalability and on-demand ammonia generation. The electrochemical catalysis community has progressed somewhat towards the fabrication of efficient electrochemical nitrogen reduction materials, but most efforts are plagued by low Faradaic efficiencies due to the competing hydrogen production reaction, which dominates all metal-based catalyst surfaces [19, 79–81]. Therefore, natural biocatalytic systems that are able to efficiently catalyze NRR reactions are highly desirable alternatives. Recently, several works on biological catalysts based on protein compounds have been published. These state that the conversion of ­N2 into ammonia occurs naturally in diazotrophic microorganisms through the enzyme nitrogenase, as shown in Fig. 6. [77] Also, they reported that nitrogenase operates at mild conditions of around 150 bar. The synthesis of N ­ H3 from dinitrogen by nitrogenase follows this reaction under optimal conditions.

N2 + 8H+ + 16MgATP + 8e− → 2NH3 + H2 + 16MgADP + 16Pi (where ATP is adenosine tri-phosphate, ADP is adenosine diphosphate and Pi is inorganic phosphate). The reaction includes the obligatory hydrolysis of ATP to release stored chemical energy and kinetic limitations of nitrogen reduction. According to the results of this work, for the molecule of nitrogen that is reduced, two molecules of ammonia are produced, and protons are also reduced to form one molecule of hydrogen. Following these ideas, theoretical scientists are starting to work on the design of nature-inspired NRR nanostructured materials. In this regard, a giant advance was recently achieved by He and coworkers in this innovative research line [82]. They reported the first theoretically designed asymmetrical dual-metal dimer catalytic centers embedded on N-doped carbon nanostructures toward the electrochemical reduction of ­N2–NH3 inspired by the function of FeMo cofactors in the nitrogenase molecules. The simulated Mo–Ru, Mo–Co, Mo–W, Mo–Fe and Fe–Ru dimers exhibited ultra-low onset potentials of only 0.17, 0.27, 0.28, 0.36 and 0.39 V vs reversible hydrogen electrode (RHE), opening the way for the development of promising nature-inspired NRR electrocatalysts (Fig. 7) [82].

6 Conclusions The development of protein-based nanomaterials for water-splitting reactions has becoming cutting-edge research in recent years. Their unique properties to competently catalyze several electrocatalytic reactions at the nanometric scale make them excellent candidates for the fabrication of promising enzymatic biofuel cells. Undoubtedly, the role of proteins as effective nanomachines in the overall catalytic process is crucial. In this direction, it is well-established that conformational and/or structural changes in protein structures significantly enhance their

13

250

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





Fig. 6  Nitrogenase enzyme structure and functions. a Diagram of one half of the nitrogenase complex and electron transfer. b FeMo cofactor and its environment (reprinted from Ref. [77])

electrocatalytic yields. Despite these achievements, knowledge of the mechanisms that govern electrocatalytic reactions in nanobiosystems as well as the synergistic behavior between enzymes and nanosized materials are still in their infancy. Therefore, a lot of effort should be delivered to shed light on the underlying biochemistry of protein-based nanoelectrocatalysts. We envision that this upcoming Reprinted from the journal

251

13





Topics in Current Chemistry (2020) 378:43

Fig. 7  The calculated Gibbs free energy changes for the first and last hydrogenation steps of nitrogen reduction reaction (NRR) on different MN4-NG and M1M2N6-NG catalysts (reprinted from Ref. [82])

knowledge will enable the design of biocatalytic nanohybrids with unprecedented electrocatalytic properties. Funding  This work was funded by the Prof. Rafael Luque Grant number CTQ2016-78289-P.

References 1. Kornienko N, Zhang JZ, Sakimoto KK, Yang PD, Reisner E (2018) Nat Nanotechnol 13:890. https​ ://doi.org/10.1038/s4156​5-018-0251-7 2. Kornienko N, Ly KH, Robinson WE, Heidary N, Zhang JZ, Reisner E (2019) Acc Chem Res 52:1439. https​://doi.org/10.1021/acs.accou​nts.9b000​87 3. Hitaishi VP, Mazurenko I, Harb M, Clement R, Taris M, Castano S, Duche D, Lecomte S, Ilbert M, de Poulpiquet A, Lojou E (2018) ACS Catal 8:12004. https​://doi.org/10.1021/acsca​tal.8b034​43 4. Lancaster L, Abdallah W, Banta S, Wheeldon I (2018) Chem Soc Rev 47:5177. https​://doi. org/10.1039/c8cs0​0085a​ 5. Xiao XX, Xia HQ, Wu RR, Bai L, Yan L, Magner E, Cosnier S, Lojou E, Zhu ZG, Liu AH (2019) Chem Rev 119:9509. https​://doi.org/10.1021/acs.chemr​ev.9b001​15 6. Franco A, Cano M, Giner-Casares JJ, Rodriguez-Castellon E, Luque R, Puente-Santiago AR (2019) Chem Commun 55:4671. https​://doi.org/10.1039/c9cc0​1625b​ 7. Patra S, Sene S, Mousty C, Serre C, Chausse A, Legrand L, Steunou N (2016) ACS Appl Mater Interfaces 8:20012. https​://doi.org/10.1021/acsam​i.6b052​89 8. Le Goff A, Holzinger M, Cosnier S (2015) Cell Mol Life Sci 72:941. https​://doi.org/10.1007/s0001​ 8-014-1828-4 9. Lalaoui N, Rousselot-Pailley P, Robert V, Mekmouche Y, Villalonga R, Holzinger M, Cosnier S, Tron T, Le Goff A (2016) ACS Catal 6:1894. https​://doi.org/10.1021/acsca​tal.5b024​42 10. Campbell E, Kaltenbach M, Correy GJ, Carr PD, Porebski BT, Livingstone EK, Afriat-Jurnou L, Buckle AM, Weik M, Hollfelder F, Tokuriki N, Jackson CJ (2016) Nat Chem Biol 12:944. https​:// doi.org/10.1038/nchem​bio.2175 11. Singh P, Francis K, Kohen A (2015) ACS Catal 5:3067. https​://doi.org/10.1021/acsca​tal.5b003​31

13

252

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





12. Otten R, Liu L, Kenner LR, Clarkson MW, Mavor D, Tawfik DS, Kern D, Fraser JS (2018) Nat Commun 9:1314. https​://doi.org/10.1038/s4146​7-018-03562​-9 13. Ma BY, Nussinov R (2016) Nat Chem Biol 12:890. https​://doi.org/10.1038/nchem​bio.2212 14. Hanoian P, Liu CT, Hammes-Schiffer S, Benkovic S (2015) Acc Chem Res 48:482. https​://doi. org/10.1021/ar500​390e 15. Stiller JB, Kerns SJ, Hoemberger M, Cho YJ, Otten R, Hagan MF, Kern D (2019) Nat Catal 2:726. https​://doi.org/10.1038/s4192​9-019-0307-6 16. Wu F, Su L, Yu P, Mao LQ (2017) J Am Chem Soc 139:1565. https​://doi.org/10.1021/jacs.6b114​69 17. Lalaoui N, David R, Jamet H, Holzinger M, Le Goff A, Cosnier S (2016) ACS Catal 6:4259. https​:// doi.org/10.1021/acsca​tal.6b007​97 18. Bollella P, Hibino Y, Kano K, Gorton L, Antiochia R (2018) ACS Catal 8:10279. https​://doi. org/10.1021/acsca​tal.8b027​29 19. Puente-Santiago AR, Rodriguez-Padron D, Quan XB, Batista MJM, Martins LO, Verrna S, Varma RS, Zhou J, Luque R (2019) ACS Sustain Chem Eng 7:1474. https​://doi.org/10.1021/acssu​schem​ eng.8b051​07 20. Le Goff A, Holzinger M (2018) Sustain Energy Fuels 2:2555. https​://doi.org/10.1039/c8se0​0374b​ 21. Al-Lolage FA, Bartlett PN, Gounel S, Staigre P, Mano N (2019) ACS Catal 9:2068. https​://doi. org/10.1021/acsca​tal.8b043​40 22. Dagys M, Laurynenas A, Ratautas D, Kulys J, Vidziunaite R, Talaikis M, Niaura G, Marcinkeviciene L, Meskys R, Shleev S (2017) Energy Environ Sci 10:498. https​://doi.org/10.1039/c6ee0​ 2232d​ 2 3. Lalaoui N, Le Goff A, Holzinger M, Mermoux M, Cosnier S (2015) Chem Eur J 21:3198. https​:// doi.org/10.1002/chem.20140​5557 24. Navaee A, Salimi A (2015) J Mat Chem A 3:7623. https​://doi.org/10.1039/c4ta0​6867j​ 25. Ben Tahar A, Romdhane A, Lalaoui N, Reverdy-Bruas N, Le Goff A, Holzinger M, Cosnier S, Chaussy D, Belgacem N (2018) J Power Sources 408:1. https​://doi.org/10.1016/j.jpows​ our.2018.10.059 26. Xu R, Tang RZ, Zhou QJ, Li FT, Zhang BR (2015) Chem Eng J 262:88. https​://doi.org/10.1016/j. cej.2014.09.072 27. Guo DH, Shibuya R, Akiba C, Saji S, Kondo T, Nakamura J (2016) Science 351:361. https​://doi. org/10.1126/scien​ce.aad08​32 28. Singh SK, Takeyasu K, Nakamura J (2019) Adv Mater 31:e1804297. https​://doi.org/10.1002/ adma.20180​4297 29. Sokolov SV, Sepunaru L, Compton RG (2017) Appl Mater Today 7:82. https​://doi.org/10.1016/j. apmt.2017.01.005 30. Page CC, Moser CC, Dutton PL (2003) Curr Opin Chem Biol 7:551. https​://doi.org/10.1016/j. cbpa.2003.08.005 31. Sjodin M, Styring S, Wolpher H, Xu YH, Sun LC, Hammarstrom L (2005) J Am Chem Soc 127:3855. https​://doi.org/10.1021/ja044​395o 32. Butterfield CN, Tao LZ, Chacon KN, Spiro TG, Blackburn NJ, Casey WH, Britt RD, Tebo BM (2015) BBA-Proteins Proteom 1854:1853. https​://doi.org/10.1016/j.bbapa​p.2015.08.012 33. Yates NDJ, Fascione MA, Parkin A (2018) Chem Eur J 24:12164. https​://doi.org/10.1002/ chem.20180​0750 34. Saboe PO, Conte E, Farell M, Bazan GC, Kumar M (2017) Energy Environ Sci 10:14. https​://doi. org/10.1039/c6ee0​2801b​ 35. Shafaat HS, Rudiger O, Ogata H, Lubitz W (2013) Biochim Biophys 1827:986. https​://doi. org/10.1016/j.bbabi​o.2013.01.015 36. Mazurenko I, Wang X, de Poulpiquet A, Lojou E (2017) Sustain Energ Fuels 1:1475. https​://doi. org/10.1039/c7se0​0180k​ 37. Gentil S, Mansor SMC, Jamet H, Cosnier S, Cavazza C, Le Goff A (2018) ACS Catal 8:3957. https​ ://doi.org/10.1021/acsca​tal.8b007​08 38. McDonald TJ, Svedruzic D, Kim YH, Blackburn JL, Zhang SB, King PW, Heben MJ (2007) Nanoletters 7:3528. https​://doi.org/10.1021/nl072​319o 39. Enguita FJ, Martins LO, Henriques AO, Carrondo MA (2003) J Biol Chem 278:19416. https​://doi. org/10.1074/jbc.M3012​51200​ 40. Gell DA (2018) Blood Cell Mol Dis 70:13. https​://doi.org/10.1016/j.bcmd.2017.10.006 41. Yeung N, Lin YW, Gao YG, Zhao X, Russell BS, Lei LY, Miner KD, Robinson H, Lu Y (2009) Nature 462:1079. https​://doi.org/10.1038/natur​e0862​0 Reprinted from the journal

253

13





Topics in Current Chemistry (2020) 378:43

42. Maia LB, Fonseca L, Moura I, Moura JJG (2016) J Am Chem Soc 138:8834. https​://doi. org/10.1021/jacs.6b039​41 43. Parimi NS, Umasankar Y, Atanassov P, Ramasamy RP (2012) Acs Catalysis 2:38. https​://doi. org/10.1021/cs200​527c 44. Jin B, Wang GX, Millo D, Hildebrandt P, Xia XH (2012) J Phys Chem C 116:13038. https​://doi. org/10.1021/jp303​740e 45. Mukherjee S, Mukherjee A, Bhagi-Damodaran A, Mukherjee M, Lu Y, Dey A (2015) Nat Commun: 6:8467. https​://doi.org/10.1038/ncomm​s9467​ 46. Liu JW, Zheng Y, Hong ZL, Cai K, Zhao F, Han HY (2016) Sci Adv 2:144. https​://doi.org/10.1126/ sciad​v.16008​58 47. Hou YN, Liu H, Han JL, Cai WW, Zhou JZ, Wang AJ, Cheng HY (2016) ACS Sustain Chem Eng 4:5392. https​://doi.org/10.1021/acssu​schem​eng.6b006​47 48. Reguera G, McCarthy KD, Mehta T, Nicoll JS, Tuominen MT, Lovley DR (2005) Nature 435:1098. https​://doi.org/10.1038/natur​e0366​1 49. Sekretaryova AN, Vagin MY, Turner APF, Eriksson M (2016) J Am Chem Soc 138:2504. https​:// doi.org/10.1021/Jacs.5b131​49 50. Rodriguez-Padron D, Puente-Santiago AR, Caballero A, Balu AM, Romero AA, Luque R (2018) Nanoscale 10:3961. https​://doi.org/10.1039/c8nr0​0512e​ 51. Wang XQ, Li ZJ, Qu YT, Yuan TW, Wang WY, Wu Y, Li YD (2019) Chem 5:1486. https​://doi. org/10.1016/j.chemp​r.2019.03.002 52. Luo MC, Zhao ZL, Zhang YL, Sun YJ, Xing Y, Lv F, Yang Y, Zhang X, Hwang S, Qin YN, Ma JY, Lin F, Su D, Lu G, Guo SJ (2019) Nature 574:81. https​://doi.org/10.1038/s4158​6-019-1603-7 53. Dai LM (2019) Adv Mater 31:1970322. https​://doi.org/10.1002/adma.20190​0973 54. Guo CZ, Chen CG, Luo ZL (2014) J Power Sources 245:841. https​://doi.org/10.1016/j.jpows​ our.2013.07.037 55. Chatterjee K, Ashokkumar M, Gullapalli H, Gong YJ, Vajtai R, Thanikaivelan P, Ajayan PM (2018) Carbon 130:645. https​://doi.org/10.1016/j.carbo​n.2018.01.052 56. Franco A, Cebrian-Garcia S, Rodriguez-Padron D, Puente-Santiago AR, Munoz-Batista MJ, Caballero A, Balu AM, Romero AA, Luque R (2018) ACS Sustain Chem Eng 6:11058. https​://doi. org/10.1021/acssu​schem​eng.8b025​29 57. Guo CZ, Liao WL, Li ZB, Sun LT, Chen CG (2015) Nanoscale 7:15990. https​://doi.org/10.1039/ c5nr0​3828f​ 5 8. Firpo V, Le JM, Pavone V, Lombardi A, Bren KL (2018) Chem Sci 9:8582. https​://doi.org/10.1039/ c8sc0​1948g​ 5 9. Rodriguez-Padron D, Puente-Santiago AR, Cano M, Caballero A, Munoz-Batista MJ, Luque R (2020) ACS Appl Mater Interf 12:2207. https​://doi.org/10.1021/acsam​i.9b135​71 60. Kaila VRI, Verkhovsky MI, Wikstrom M (2010) Chem Rev 110:7062. https​://doi.org/10.1021/cr100​ 2003 61. Wikstrom M, Sharma V, Kaila VRI, Hosler JP, Hummer G (2015) Chem Rev 115:2196. https​://doi. org/10.1021/cr500​448t 62. Gorlin Y, Jaramillo TF (2010) J Am Chem Soc 132:13612. https​://doi.org/10.1021/ja104​587v 63. Lee K, Zhang L, Lui H, Hui R, Shi Z, Zhang JJ (2009) Electrochim Acta 54(4704):44. https​://doi. org/10.1016/j.elect​acta.2009.03.081 64. Dey S, Mondal B, Chatterjee S, Rana A, Amanullah SK, Dey A (2017) Nat Rev Chem 1:15. https​:// doi.org/10.1038/s4157​0-017-0098 65. Burgess JD, Rhoten MC, Hawkridge FM (1998) Langmuir 14:2467. https​://doi.org/10.1021/la971​ 1995 66. Yoshikawa S, Shimada A (2015) Chem Rev 115:1936. https​://doi.org/10.1021/cr500​266a 67. Cracknell JA, Blanford CF (2012) Chem Sci 3:1567. https​://doi.org/10.1039/c2sc0​0632d​ 68. Blanford CF, Heath RS, Armstrong FA (2007) Chem Commun 22:1710. https​://doi.org/10.1039/ b7031​14a 69. Jones SM, Solomon EI (2015) Cell Mol Life Sci 72:869. https​://doi.org/10.1007/s0001​8-014-1826-6 70. Collman JP, Rapta M, Broring M, Raptova L, Schwenninger R, Boitrel B, Fu L, L’Her M (1999) J Am Chem Soc 121:1387. https​://doi.org/10.1021/ja983​351a 71. Collman JP, Devaraj NK, Decreau RA, Yang Y, Yan YL, Ebina W, Eberspacher TA, Chidsey CED (2007) Science 315:1565. https​://doi.org/10.1126/scien​ce.11358​44 72. Shi CN, Anson FC (1990) Inorg Chem 29:4298. https​://doi.org/10.1021/ic003​46a02​7

13

254

Reprinted from the journal

Topics in Current Chemistry (2020) 378:43





73. Rigsby ML, Wasylenko DJ, Pegis ML, Mayer JM (2015) J Am Chem Soc 137:4296. https​://doi. org/10.1021/jacs.5b003​59 74. Collman JP, Denisevich P, Konai Y, Marrocco M, Koval C, Anson FC (1980) J Am Chem Soc 102:6027. https​://doi.org/10.1021/ja005​39a00​9 75. Alba-Molina D, Rodriguez-Padron D, Puente-Santiago AR, Giner-Casares JJ, Martin-Romero MT, Camacho L, Martins LO, Munoz-Batista MJ, Cano M, Luque R (2019) Nanoscale 11:1549. https​:// doi.org/10.1039/c8nr0​6001k​ 76. Montoya JH, Tsai C, Vojvodic A, Norskov JK (2015) Chemsuschem 8:2180. https​://doi.org/10.1002/ cssc.20150​0322 77. Foster SL, Bakovic SIP, Duda RD, Maheshwari S, Milton RD, Minteer SD, Janik MJ, Renner JN, Greenlee LF (2018) Nat Catal 1:490. https​://doi.org/10.1038/s4192​9-018-0092-7 78. Licht S, Cui BC, Wang BH, Li FF, Lau J, Liu SZ (2014) Science 345:637. https​://doi.org/10.1126/ scien​ce.12542​34 79. Zhou FL, Azofra LM, Ali M, Kar M, Simonov AN, McDonnell-Worth C, Sun CH, Zhang XY, MacFarlane DR (2017) Energy Environ Sci 10:2516. https​://doi.org/10.1039/c7ee0​2716h​ 80. Kim H, Renteria-Marquez A, Islam MD, Chavez LA, Rosales CAG, Ahsan MA, Tseng TLB, Love ND, Lin YR (2019) J Am Ceramd Soc 102:3685. https​://doi.org/10.1111/jace.16242​ 81. Dominguez N, Torres B, Barrera LA, Rincon JE, Lin YR, Chianelli RR, Ahsan MA, Noveron JC (2018) ACS Omega 3:10243. https​://doi.org/10.1021/acsom​ega.8b006​54 82. He T, Puente-Santiago AR, Du A (2020) J Catal 388:77–83.https​://doi.org/10.1016/j. jcat.2020.05.009

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Reprinted from the journal

255

13