Nothing Special   »   [go: up one dir, main page]

Nanozymes Next Wave of Artificial Enzymes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 134

SPRINGER BRIEFS IN MOLECULAR SCIENCE

Xiaoyu Wang
Wenjing Guo
Yihui Hu
Jiangjiexing Wu
Hui Wei

Nanozymes: Next
Wave of Artificial
Enzymes

12 3
SpringerBriefs in Molecular Science
More information about this series at http://www.springer.com/series/8898
Xiaoyu Wang Wenjing Guo

Yihui Hu Jiangjiexing Wu

Hui Wei

Nanozymes: Next Wave


of Artificial Enzymes

123
Xiaoyu Wang Jiangjiexing Wu
Department of Biomedical Engineering, College of Department of Biomedical Engineering, College of
Engineering and Applied Sciences, Collaborative Engineering and Applied Sciences, Collaborative
Innovation Center of Chemistry for Life Sciences, Innovation Center of Chemistry for Life Sciences,
Nanjing National Laboratory of Microstructures Nanjing National Laboratory of Microstructures
Nanjing University Nanjing University
Nanjing Nanjing
China China

Wenjing Guo Hui Wei


Department of Biomedical Engineering, College of Department of Biomedical Engineering, College of
Engineering and Applied Sciences, Collaborative Engineering and Applied Sciences, Collaborative
Innovation Center of Chemistry for Life Sciences, Innovation Center of Chemistry for Life Sciences,
Nanjing National Laboratory of Microstructures Nanjing National Laboratory of Microstructures
Nanjing University Nanjing University
Nanjing Nanjing
China China

Yihui Hu
Department of Biomedical Engineering, College of
Engineering and Applied Sciences, Collaborative
Innovation Center of Chemistry for Life Sciences,
Nanjing National Laboratory of Microstructures
Nanjing University
Nanjing
China

ISSN 2191-5407 ISSN 2191-5415 (electronic)


SpringerBriefs in Molecular Science
ISBN 978-3-662-53066-5 ISBN 978-3-662-53068-9 (eBook)
DOI 10.1007/978-3-662-53068-9

Library of Congress Control Number: 2016946947

© The Author(s) 2016


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, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer-Verlag GmbH Berlin Heidelberg
Preface

This book is intended to describe the concepts, the up-to-date developments, and
the perspectives of the field of nanozymes that has been rapidly growing over the
past decades. Nanozymes are nanomaterials with enzymatic characteristics. As one
of the most exciting fields, the research of nanozymes lies at the interface of
chemistry, biology, materials, and nanotechnology.
It is counterintuitive to use nanomaterials to mimic natural enzymes since the
two seem to be very different from each other. A careful comparison, however,
would reveal that they share many features together. For examples, both of them
have nanoscaled sizes, irregular shapes, rich surface chemistry, etc. It is these
similarities that enable nanomaterials to imitate natural enzymes.
Due to the enormous amounts of literature published in the field, it is impossible
to provide a comprehensive description of nanozymes here. Instead, it aims to
provide a broad picture of nanozymes in the context of artificial enzyme research.
Representative examples are discussed to highlight the nanomaterials with enzyme
mimicking activities, their catalytic mechanisms, and their promising applications
in various areas, ranging from biosensing and cancer diagnostics to tissue engi-
neering and therapeutics.
Chapter 1 describes the brief history of nanozymes research in the course of
natural enzymes and artificial enzymes research. It also compares nanozymes with
natural enzymes and artificial enzymes to highlight their unique characteristics.
Chapters 2–5 discuss the different nanomaterials used for mimicking various natural
enzymes, from carbon-based (Chap. 2) and metal-based (Chap. 3) nanomaterials to
metal oxide-based nanomaterials (Chap. 4) and other nanomaterials (Chap. 5). In
each of these chapters, the nanomaterials’ enzyme mimetic activities, the catalytic
mechanisms, and the key applications are covered. In Chap. 6, the current chal-
lenges and future directions of nanozymes research are summarized, which if
achieved will help to fulfill the great potentials of nanozymes.
The purpose of this book is not only to provide insightful knowledge of nano-
zymes but also to attract more researchers into the field and to inspire them to
further broaden the field. Due to the importance of nanozymes and professional

v
vi Preface

writing with plenty of color illustrations and tables, this book should be an ideal
choice for readers from different areas, such as chemistry, materials, nanoscience
and nanotechnology, biomedical and clinical studies, environment, green chem-
istry, novel catalysts, etc.
I wish to express my appreciation to all the excellent scholars around the world
who have contributed and will continue to contribute to the fields of nanozymes.
I would also like to thank my lab members and my collaborators for their contri-
butions to this exciting field. I thank my advisors Profs. Erkang Wang, Xinghua
Xia, Yi Lu, and Shuming Nie for their guidance, support, and encouragement. I am
much indebted to June Tang for her patience during the writing of this book.
I thank Nanjing University, National Natural Science Foundation of China, 973
Program, Natural Science Foundation of Jiangsu Province, Shuangchuang Program
of Jiangsu Province, PAPD program, Fundamental Research Funds for Central
Universities, Six Talents Summit Program of Jiangsu Province, Open Funds of the
State Key Laboratory of Electroanalytical Chemistry, Open Funds of the State Key
Laboratory of Analytical Chemistry for Life Science, Open Funds of the State Key
Laboratory for Chemo/Biosensing and Chemometrics, and Thousand Talents
Program for Young Researchers for providing the academic environment and the
financial support of our research.

Nanjing Hui Wei


April 2016
Contents

1 Introduction to Nanozymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2 Carbon-Based Nanomaterials for Nanozymes . . . . . . . . . . . . . . . . . . . 7
2.1 Fullerene and Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1 Fullerene and Derivatives as Nuclease Mimics . . . . . . . . . . 8
2.1.2 Fullerene and Derivatives as SOD Mimics . . . . . . . . . . . . . 9
2.1.3 Fullerene Derivatives as Peroxidase Mimics . . . . . . . . . . . . 12
2.2 Graphene and Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.1 Graphene and Its Derivatives as Peroxidase Mimics . . . . . . 12
2.2.2 Decorated Graphene (or Its Derivatives)
as Peroxidase Mimics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Carbon Nanotubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3.1 Carbon Nanotubes as Peroxidase Mimics . . . . . . . . . . . . . . 20
2.3.2 Carbon Nanotubes as Other Enzyme Mimics . . . . . . . . . . . 22
2.4 Other Carbon-Based Nanomaterials . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4.1 Other Carbon Nanomaterials as Peroxidase Mimics . . . . . . 24
2.4.2 Other Carbon Nanomaterials as SOD Mimics . . . . . . . . . . . 24
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3 Metal-Based Nanomaterials for Nanozymes . . . . . . . . . . . . . . . . .... 31
3.1 Metal Nanomaterials with Catalytic Monolayers (Type I). . . . .... 31
3.1.1 AuNPs Protected by Alkanethiol with Catalytic
Terminal Moieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 32
3.1.2 AuNPs Protected by Alkanethiol with Non-covalently
Assembled Catalytic Moieties . . . . . . . . . . . . . . . . . . . .... 37
3.1.3 AuNPs Protected by Thiolated Biomolecules . . . . . . . .... 39

vii
viii Contents

3.2 Metal Nanomaterials with Intrinsic Enzyme Mimicking


Activities (Type II) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2.1 Metal Nanomaterials as GOx Mimics . . . . . . . . . . . . . . . . . 40
3.2.2 Metal Nanomaterials as Multiple Enzyme Mimics . . . . . . . 41
3.2.3 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4 Metal Oxide-Based Nanomaterials for Nanozymes . . . . . . . . . . . . . . . 57
4.1 Cerium Oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.1.1 Cerium Oxide as SOD Mimics . . . . . . . . . . . . . . . . . . . . . . 58
4.1.2 Cerium Oxide as Catalase Mimics . . . . . . . . . . . . . . . . . . . 64
4.1.3 Cerium Oxide as Peroxidase Mimics . . . . . . . . . . . . . . . . . 66
4.1.4 Cerium Oxide as Oxidase Mimics . . . . . . . . . . . . . . . . . . . . 66
4.1.5 Cerium Oxide as Other Mimics . . . . . . . . . . . . . . . . . . . . . 67
4.2 Iron Oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.2.1 Iron Oxide as Peroxidase Mimics . . . . . . . . . . . . . . . . . . . . 68
4.2.2 Iron Oxide as Other Enzyme Mimics . . . . . . . . . . . . . . . . . 76
4.3 Other Metal Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.3.1 Vanadium Oxide as Enzyme Mimics . . . . . . . . . . . . . . . . . 78
4.3.2 Cobalt Oxide as Enzyme Mimics . . . . . . . . . . . . . . . . . . . . 78
4.3.3 Copper Oxide as Enzyme Mimics . . . . . . . . . . . . . . . . . . . . 81
4.3.4 MoO3, TiO2, MnO2, RuO2 as Enzyme Mimics . . . . . . . . . . 81
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5 Other Nanomaterials for Nanozymes . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.1 Prussian Blue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.2 Metal-Organic Frameworks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.3 Metal Chalcogenides. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.4 Metal Hydroxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.5 Miscellaneous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6 Challenges and Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Abbreviations

4-AAP 4-aminoantipyrine
ABTS 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)
AgNP Silver nanoparticle
AuNC Gold nanocluster
AuNP Gold nanoparticle
BA Benzoic acid
BSA Bovine serum albumin
CEA Carcinoembryonic antigen
CNT Carbon nanotube
Color. Colorimetric
DAB Diazoaminobenzene
DOPA Dopamine
DPD N,N-diethyl-p-phenylenediamine sulfate
dsDNA Double-stranded DNA
E-chem Electrochemical
ELISA Enzyme-linked immunosorbent assay
EPR Electron paramagnetic resonance
Fluor. Fluorometric
HPNP 2-hydroxypropyl-4-nitrophenylphosphate
HRP Horseradish peroxidase
LDH Layered double hydroxide
LOD Limit of detection
Meth Methods
MNPs Magnetic nanoparticles
NMDA N-methyl-D-aspartate
NPs Nanoparticles
OPD o-phenylenediamine
PDDA Poly(diallyldimethylammonium chloride)
PLGA Poly(D,L-lactic-co-glycolic acid)
PMIDA N-(phosphonomethyl)iminodiacetic acid

ix
x Abbreviations

PSA Prostate-specific antigen


PSS Poly(styrenesulfonate)
PVDF Polyvinylidene difluoride
Ref References
SBA-15 Santa Barbara Amorphous type material
SOD Superoxide dismutase
ssDNA Single-stranded DNA
TMB 3,3′,5,5′-tetramethylbenzidine
Chapter 1
Introduction to Nanozymes

Abstract Natural enzymes play vital roles in biological reactions in living systems.
However, some intrinsic drawbacks, such as ease of denaturation, laborious
preparation, high cost, and difficulty of recycling, have limited their practical
applications. To tackle these problems, intensive efforts have been devoted to
developing natural enzymes’ alternatives called “artificial enzymes.” As an
emerging research area of artificial enzymes, nanozymes, the catalytic nanomate-
rials with enzyme-like characteristics, have attracted researchers’ enormous atten-
tions. In this chapter, after the brief description of the history of nanozymes research
in the course of natural enzymes and artificial enzymes research, a comparison
between nanozymes and natural enzymes as well as artificial enzymes is made.
Such a comparison highlights the unique characteristics of nanozymes, such as their
size-(shape-, structure-, composition-)tunable catalytic activities, large surface area
for modification and bioconjugation, multiple functions besides catalysis, smart
response to external stimuli, etc.

 
Keywords Natural enzymes Artificial enzymes Nanozymes Enzyme mim- 
ics  Catalytic nanomaterials 
Nanobiology 
Functional nanomaterials 
 
Biological catalysts Biomimetic chemistry Nanomaterials with enzyme-like
characteristics

Natural enzymes are ubiquitous biocatalysts that play central roles in virtually all
the biological reactions in living systems [1]. Since they catalyze the reactions with
remarkable efficiency and extraordinary specificity at mild conditions (such as room
temperature, ambient pressure, aqueous solutions, etc.), natural enzymes have been
extensively explored for various applications beyond living systems. For instance,
they have been widely used in biomedicine, clinic, environmental and food
industry. On the other hand, natural enzymes are proteins or ribonucleic acids,
which inevitably have several intrinsic drawbacks, such as ease of denaturation,
laborious preparation, high cost, difficulty of recycling, etc. These drawbacks have
in turn limited their practical applications.

© The Author(s) 2016 1


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_1
2 1 Introduction to Nanozymes

Fig. 1.1 A brief timeline for the development of artificial enzymes (natural enzymes are also
listed for comparison). Reprinted from Ref. [2], Copyright 2016, with permission from Royal
Society of Chemistry

To tackle these drawbacks, intensive efforts have been devoted to developing


natural enzymes’ alternatives called “artificial enzymes” (or “enzyme mimics”) since
1950s (Fig. 1.1) [3]. Artificial enzymes aim at “imitating the catalytic processes
that occur in living systems,” as defined by Breslow [3]. In their pioneering work,
Breslow and others have used cyclodextrins and their derivatives to mimic varieties
of enzymes, ranging from thiamine pyrophosphate and pyridoxal phosphate to
hydrolytic enzymes and even cytochrome P-450 [3]. Inspired by the success of
these studies, researchers have investigated numerous types of materials like metal
complexes, polymers, supramolecules, and biomolecules (such as nucleic acids,
catalytic antibodies, and proteins) for mimicking various kinds of natural enzymes
[3]. For example, synthetic polymers with enzyme-like catalytic activities have
been studied by Klotz et al. [4]. To date, enormous progress has been made in the
field of artificial enzymes (Fig. 1.1), as evidenced by the publication of numerous
excellent reviews and even several monographs on the topic [3, 5–22].
Over the past two decades, along with the remarkable achievements made in the
field of nanotechnology, varieties of functional nanomaterials have been discovered
to possess unexpected enzyme-mimicking catalytic activities (Fig. 1.1). These
emerging functional nanomaterials are now collectively termed as “nanozymes”.
The term “nanozymes” was coined by Pasquato, Scrimin, and their coworkers in
2004 to describe the gold nanoparticle-based transphosphorylation mimics resulting
from the self-assembly of triazacyclonane-functionalized thiols onto the surface of
gold nanoparticles [23]. Later, in their comprehensive review published in 2013,
Wei and Wang defined “nanozymes” as “nanomaterials with enzyme-like charac-
teristics” [24].
As a new type of promising artificial enzymes, nanozymes have attracted con-
siderable attentions, particularly in recent years. As shown in Fig. 1.2, since the
seminal work on fullerene derivatives-based DNase mimics (i.e., their capability for
1 Introduction to Nanozymes 3

Fig. 1.2 Number of published papers on nanozymes by the end of 2015. The data is based on
Web of Science (April 2016)

oxidative DNA cleavage) in the early 1990s [26], incredible growth has been
witnessed in the field of nanozymes by the exponential number of publications. By
the end of 2015, more than 740 papers on nanozymes have been published, among
which 722 were published after 2007. Recently, a special issue has been devoted to
nanozymes [27]. The growing interests in nanozymes can be attributed to their
unique characteristics over natural enzymes and even conventional artificial
enzymes. Nanozymes are unique in several aspects, such as their size- (shape-,
structure-, composition-)tunable catalytic activities, large surface area for modifi-
cation and bioconjugation, multiple functions besides catalysis, smart response to
external stimuli, etc. (Table 1.1) [24]. To date, plenty of nanomaterials have been
investigated to mimic diverse natural enzymes, which have already found many
interesting applications [2, 24, 25, 28–42].
In the following chapters (Chaps. 2–5), nanozymes are discussed based on the
nanomaterials rather than the natural enzymes which they mimic because many
nanomaterials have exhibited multiple enzyme-mimicking activities. We do not
attempt to cover all the published papers on nanozymes in this book. Instead, for
each typical nanomaterial, we mainly focus on its enzyme-mimicking activities,
catalytic mechanisms, and the key applications. In the last chapter (Chap. 6), we
discuss the current challenges and future prospects that nanozyme research is
currently facing to fulfill its great potentials.
4 1 Introduction to Nanozymes

Table 1.1 Comparison between nanozymes and othersa, b

Characteristics
Nanozymes (1)Low cost
(2)Easy for mass production
(3)Robustness to harsh environments
(4)High stability
(5)Long-term storage
(6)Tunable activity
(7)Size- (shape-, structure-, composition-) dependent
properties
(8) Multifunction
(9) Easy for further modification (such as bioconjugation)
(10) Smart response to external stimuli
(11) Self-assembly
Conventional artificial (1) Low cost
enzymes (2) Easy mass production
(3) Robustness to harsh environments
(4) High stability
(5) Long-term storage
(6) Tunable activity
(7) Established methods for preparation and characterization
(8) Uniform size and defined structures of molecular mimics
(9) Smaller size (compared with nanozymes)
Natural enzymes (1) High catalytic efficiency
(2) High substrate specificity
(3) High (enantio)selectivity
(4) Sophisticated three-dimensional structures
(5) Wide range of catalytic reactions
(6) Tunable activity
(7) Good biocompatibility
(8) Rational design via protein engineering and computation
a
The items in italic font are unique for nanozymes compared with conventional artificial enzymes
b
The table was adapted from Ref. [24], Copyright 2013, with permission from Royal Society of
Chemistry

References

1. Nelson, D. L., & Cox, M. M. (2008). Lehninger principles of biochemistry (5th ed., pp. 183–
229). New York: W. H. Freeman and Company.
2. Wang, X. Y., Hu, Y. H., & Wei, H. (2016). Nanozymes in bionanotechnology: From sensing
to therapeutics and beyond. Inorganic Chemistry Frontiers, 3, 41–60.
3. Breslow, R. (2005). Artificial enzymes. Weinheim: Wiley-VCH.
4. Klotz, I. M., Royer, G. P., & Scarpa, I. S. (1971). Synthetic derivatives of polyethyleneimine
with enzyme-like catalytic activity (synzymes). Proceedings of the National Academy of
Sciences of the United States of America, 68, 263–264.
5. Breslow, R. (1995). Biomimetic chemistry and artificial enzymes: catalysis by design.
Accounts of Chemical Research, 28, 146–153.
6. Murakami, Y., Kikuchi, J., Hisaeda, Y., & Hayashida, O. (1996). Artificial enzymes.
Chemical Reviews, 96, 721–758.
References 5

7. Breslow, R., & Dong, S. D. (1998). Biomimetic reactions catalyzed by cyclodextrins and their
derivatives. Chemical Reviews, 98, 1997–2011.
8. Kirby, A. J., Hollfelder, F. (2009). Fom enzyme models to model enzymes. The Royal Society
of Chemistry, Thomas Graham House, Science Park, Milton Road, Cambridge CB4 0WF, UK.
9. Aiba, Y., Sumaoka, J., & Komiyama, M. (2011). Artificial DNA cutters for DNA
manipulation and genome engineering. Chemical Society Reviews, 40, 5657–5668.
10. Wulff, G., & Liu, J. Q. (2012). Design of biomimetic catalysts by molecular imprinting in
synthetic polymers: The role of transition state stabilization. Accounts of Chemical Research,
45, 239–247.
11. Salvemini, D., Riley, D. P., & Cuzzocrea, S. (2002). SOD mimetics are coming of age. Nature
Reviews Drug Discovery, 1, 367–374.
12. Huang, X., Liu, X. M., Luo, Q. A., Liu, J. Q., & Shen, J. C. (2011). Artificial selenoenzymes:
Designed and redesigned. Chemical Society Reviews, 40, 1171–1184.
13. Bhabak, K. P., & Mugesh, G. (2010). Functional mimics of glutathione peroxidase:
Bioinspired synthetic antioxidants. Accounts of Chemical Research, 43, 1408–1419.
14. Friedle, S., Reisner, E., & Lippard, S. J. (2010). Current challenges of modeling diiron enzyme
active sites for dioxygen activation by biomimetic synthetic complexes. Chemical Society
Reviews, 39, 2768–2779.
15. Darbre, T., & Reymond, J. L. (2006). Peptide dendrimers as artificial enzymes, receptors, and
drug-delivery agents. Accounts of Chemical Research, 39, 925–934.
16. Molenveld, P., Engbersen, J. F. J., & Reinhoudt, D. N. (2000). Dinuclear
metallo-phosphodiesterase models: Application of calix[4]arenes as molecular scaffolds.
Chemical Society Reviews, 29, 75–86.
17. Feiters, M. C., Rowan, A. E., & Nolte, R. J. M. (2000). From simple to supramolecular
cytochrome P450 mimics. Chemical Society Reviews, 29, 375–384.
18. Riley, D. P. (1999). Functional mimics of superoxide dismutase enzymes as therapeutic
agents. Chemical Reviews, 99, 2573–2587.
19. Mertes, M. P., & Mertes, K. B. (1990). Polyammonium macrocycles as catalysts for
phosphoryl transfer—the evolution of an enzyme mimic. Accounts of Chemical Research, 23,
413–418.
20. Gloaguen, F., & Rauchfuss, T. B. (2009). Small molecule mimics of hydrogenases: Hydrides
and redox. Chemical Society Reviews, 38, 100–108.
21. Lu, Y., Yeung, N., Sieracki, N., & Marshall, N. M. (2009). Design of functional
metalloproteins. Nature, 460, 855–862.
22. Thomas, C. M., & Ward, T. R. (2005). Artificial metalloenzymes: Proteins as hosts for
enantioselective catalysis. Chemical Society Reviews, 34, 337–346.
23. Manea, F., Houillon, F. B., Pasquato, L., & Scrimin, P. (2004). Nanozymes:
Gold-nanoparticle-based transphosphorylation catalysts. Angewandte Chemie-International
Edition, 43, 6165–6169.
24. Wei, H., & Wang, E. K. (2013). Nanomaterials with enzyme-like characteristics (nanozymes):
Next-generation artificial enzymes. Chemical Society Reviews, 42, 6060–6093.
25. Bitner, B. R., Marcano, D. C., Berlin, J. M., Fabian, R. H., Cherian, L., Culver, J. C., et al.
(2012). Antioxidant carbon particles improve cerebrovascular dysfunction following traumatic
brain injury. ACS Nano, 6, 8007–8014.
26. Tokuyama, H., Yamago, S., Nakamura, E., Shiraki, T., & Sugiura, Y. (1993). Photoinduced
biochemical-activity of fullerene carboxylic-acid. Journal of the American Chemical Society,
115, 7918–7919.
27. http://www.mdpi.com/journal/molecules/special_issues/nanozymes.
28. Gao, L.-Z., & Yan, X.-Y. (2013). Discovery and current application of nanozyme. Progress in
Biochemistry and Biophysics, 40, 892–902.
29. Pasquato, L., Pengo, P., & Scrimin, P. (2005). Nanozymes: Functional nanoparticle-based
catalysts. Supramolecular Chemistry, 17, 163–171.
30. Nakamura, E., & Isobe, H. (2003). Functionalized fullerenes in water. The first 10 years of
their chemistry, biology, and nanoscience. Accounts of Chemical Research, 36, 807–815.
6 1 Introduction to Nanozymes

31. Prins, L. J. (2015). Emergence of complex chemistry on an organic monolayer. Accounts of


Chemical Research, 48, 1920–1928.
32. Xie, J. X., Zhang, X. D., Wang, H., Zheng, H. Z., & Huang, Y. M. (2012). Analytical and
environmental applications of nanoparticles as enzyme mimetics. TrAC-Trends in Analytical
Chemistry, 39, 114–129.
33. Celardo, I., Pedersen, J. Z., Traversa, E., & Ghibelli, L. (2011). Pharmacological potential of
cerium oxide nanoparticles. Nanoscale, 3, 1411–1420.
34. Lin, Y. H., Ren, J. S., & Qu, X. G. (2014). Catalytically active nanomaterials: a promising
candidate for artificial enzymes. Accounts of Chemical Research, 47, 1097–1105.
35. Karakoti, A., Singh, S., Dowding, J. M., Seal, S., & Self, W. T. (2010). Redox-active radical
scavenging nanomaterials. Chemical Society Reviews, 39, 4422–4432.
36. Lin, Y. H., Ren, J. S., & Qu, X. G. (2014). Nano-gold as artificial enzymes: hidden talents.
Advanced Materials, 26, 4200–4217.
37. He, W. W., Wamer, W., Xia, Q. S., Yin, J. J., & Fu, P. P. (2014). Enzyme-like activity of
nanomaterials. Journal of Environmental Science and Health Part C-Environmental
Carcinogenesis & Ecotoxicology Reviews, 32, 186–211.
38. Xu, C., & Qu, X. G. (2014). Cerium oxide nanoparticle: A remarkably versatile rare earth
nanomaterial for biological applications. NPG Asia Materials, 6, e60.
39. Luo, C., Li, Y., & Long, J. (2015). Recent advances in applications of nanoparticles as enzyme
mimetics. Scientia Sinica Chimica, 45, 1026–1041.
40. Gao, L. Z., & Yan, X. Y. (2016). Nanozymes: an emerging field bridging nanotechnology and
biology. Science China Life Sciences, 59, 400–402.
41. Zheng, L., Zhao, J., Niu, X., & Yang, Y. (2015). Nanomaterial-based peroxidase enzyme
mimics with applications to colorimetric analysis and electrochemical sensor. Materials
Review, 29(115–120), 129.
42. Cheng, H. J., Wang, X. Y., Wei, H. (2016) Encyclopedia of Physical Organic Chemistry
(Chapter 16), Wiley-VCH: Weinheim.
Chapter 2
Carbon-Based Nanomaterials
for Nanozymes

Abstract Carbon-based nanomaterials, such as fullerene, graphene, carbon nan-


otubes, and their derivatives, have been extensively studied to mimic various nat-
ural enzymes owing to their fascinating catalytic activities. In this chapter, their
enzyme mimetic activities (such as nuclease mimics, superoxide dismutase mimics,
peroxidase mimics, etc.) are discussed. The catalytic mechanisms are also discussed
if they have been elucidated. Representative examples for applications, from
biosensing to therapeutics, are covered.

 
Keywords Nanozymes Artificial enzymes Enzyme mimics Carbon-based 
 
nanomaterials Fullerene and derivatives Graphene and derivatives Carbon 
 
nanotubes Peroxidase mimics Superoxide dismutase mimics Nuclease mimics

Carbon-based nanomaterials, such as fullerene, carbon nanotubes (CNTs), gra-


phene, and their derivatives, have found broad applications in many areas. Owing to
their interesting catalytic activities, they have been extensively studied to mimic
various natural enzymes. In this chapter, we will discuss the nanozymes based on
these carbon nanomaterials.

2.1 Fullerene and Derivatives

Fullerene and its derivatives were among the first nanomaterials that have been
explored for mimicking natural enzymes [1, 2]. In the early 1990s, the light-induced
oxidative DNA cleavage with a C60 derivative (i.e., C60-1) was studied, indicating
that fullerene could mimic natural nuclease [1]. A few years later, the superoxide
dismutase (SOD) mimicking activity of derivatized C60 was investigated, which led
to the long-lasting research interests in fullerene-based nanozymes till today.
Fullerene and its derivatives have also been used to mimic other enzymes besides
nuclease and SOD.

© The Author(s) 2016 7


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_2
8 2 Carbon-Based Nanomaterials for Nanozymes

2.1.1 Fullerene and Derivatives as Nuclease Mimics

Nuclease catalyzes the cleavage of phosphodiester bond between two nucleotides in


a nucleic acid. Pristine fullerenes including C60 are not water soluble, which makes
them impossible to mimic enzymes in aqueous solution. Therefore, fullerenes have
been solubilized by introducing hydrophilic moieties. Nakamura group have made
water-soluble C60-1 and studied its photoinduced biochemical activities (Fig. 2.1)
[1]. Interestingly, they established that the fullerene carboxylic acid (i.e., C60-1)
oxidatively cleaved DNA under light irradiation. Since C60-1 did not bind to the
DNA to be cleaved, the cleavage was random. To address this issue, Hélène,
Nakamura and coworkers synthesized C60-2, which had a 14-mer DNA sequence

Fig. 2.1 Fullerene derivatives as nuclease mimics. a Light-induced cleavage of DNA with the
fullerene derivative C60-1. b Selective cleavage of DNA by forming a triplex with the fullerene
DNA conjugate C60-2. Adapted from Ref. [2], Copyright 2003, with permission from American
Chemical Society
2.1 Fullerene and Derivatives 9

complementary to the target DNA (Fig. 2.1) [3]. By forming a triplex, the selective
cleavage at guanine-rich sites was achieved. There are other ways to make the
selective cleavage possible. For instance, by conjugating fullerenes with DNA
intercalators (such as acridine), the formed fullerene derivative could enhance DNA
cleavage activity compared with the parent fullerene [4]. These early studies have
established that water-soluble fullerene derivatives could mimic nuclease.

2.1.2 Fullerene and Derivatives as SOD Mimics

(a) Fullerenes as SOD mimics: in vitro activities and mechanisms


Reactive oxygen species (ROS) plays both beneficial and harmful roles in living
systems. Superoxide anion, one of ROS, could cause tissue injury and associated
inflammation if it were not properly regulated. In nature, SOD has been evolved to
catalyze the disputation of superoxide anions into hydrogen peroxide and molecular
oxygen and thus protect living systems from the superoxide anion-induced damage.
To overcome the limits of natural SOD (such as limited stability and high cost),
great efforts have been devoted to developing SOD mimics. The SOD-mimicking
activities of fullerenes have been established by the seminal work from Choi and
coworkers and have since been extensively studied [5].
Inspired by the early discovery that C60 could act as a radical sponge, Choi et al.
studied the neuroprotective activities of two polyhydroxylated C60 (i.e., C60(OH)12
and C60(OH)nOm, n = 18–20, m = 3–7 hemiketal groups) [5, 7]. Surprisingly, both
of the two fullerene derivatives could scavenge free radicals and thus reduce
excitotoxic and apoptotic death of cultured cortical neurons. Later, they identified
that C60[C(COOH)2]3 with C3 symmetry (C60-C3) was more effective toward
neuron protection [8]. Since superoxide anion could be eliminated via either a
stoichiometric scavenging mechanism or a SOD-like catalytic mechanism, sys-
tematic studies including electron paramagnetic resonance (EPR) were carried out
to confirm the SOD-mimicking activity of C60-C3 [6]. The possible stoichiometric
scavenging mechanism was ruled out due to the following facts: first, no structural
modifications to C60-C3 were observed; second, the production of oxygen and
hydrogen peroxide was detected; and the absence of EPR active (paramagnetic)
products. By combining the experimental results with computational data, a cat-
alytic mechanism for SOD-like activity of C60-C3 was proposed (Fig. 2.2). The
proposed mechanism was supported by other studies using dendritic C60 derivatives
and other computational studies [9, 10].
The neuroprotective effects of fullerene-based SOD mimics were also studied
using several other cell lines. For instance, methionine-modified C60 could protect
SH-SY5Y neuroblastoma cells from lead ions (Pb2+) induced oxidative damage and
thus improved the cell survival [11].
Though water soluble fullerenes are usually required for enzyme-mimicking
studies, it has been demonstrated that pristine C60 aqueous suspension was not only
10 2 Carbon-Based Nanomaterials for Nanozymes

Fig. 2.2 Derivatized C60 as SOD mimics. a C60-C3 as a SOD mimic for catalytically converting
superoxide anion into hydrogen peroxide, water, and hydroxyl. b The proposed catalytic
mechanism. Adapted from Ref. [6], Copyright 2004, with permission from Elsevier

biocompatible but also showed protective effects on liver injury [12]. Besides
derivatization, fullerenes could also be solubilized by using the similar strategies for
hydrophobic drug solubilization. For instance, pristine C60 has been solubilized in
olive oil for various applications [13].
2.1 Fullerene and Derivatives 11

(b) Fullerenes as SOD mimics: in vivo applications


To conclusively demonstrate that C60-C3 could work in living system as a SOD
mimic, Dugan et al. employed SOD2 knockout mice as an in vivo model to
investigate the therapeutic efficacy of C60-C3 (Fig. 2.3) [6]. The SOD2 knockout
mice would die in utero or within a few days after birth owing to mitochondria
damage by oxidative species. Therefore, they were suitable models to evaluate the
in vivo effects of SOD mimics. It showed that the life span of the SOD2 defective
mice could be extended by 300 % when C60-C3 was administrated both in utero
and postnatally, demonstrating the C60-C3 could be functional alternatives to SOD2
in the studied mice. SOD2 is a manganese SOD localized in the mitochondria.
Further immunostaining indeed revealed that C60-C3 was uptaken and localized to
mitochondria [6].
These results indicated that fullerenes with SOD-mimicking activities (such as
C60-C3) may hold translational promise for treating several diseases in future.

Fig. 2.3 a Treatment of SOD2 knockout mice with C60-C3. b Percentage of Sod2−/− pups born to
Sod2+/− parents. Pregnant dams were given C60-C3 in their drinking water starting at Day 14–15 of
pregnancy. Control dams received dilute red food coloring. The percentages of Sod2−/− pups, per
litter, born to control versus C60-C3-treated dams were, respectively, 6 ± 2 % (controls, 11 L) and
20 % ± 2 (C3, 9 L) (mean ± standard error of mean, with p = 0.03 by t test, and p = 0.04 by the
nonparametric Wilcoxon rank sum test; the theoretical expected percentage is 25 %), indicating in
utero rescue of some Sod2−/− embryos. c Survival (days) of Sod2−/− pups treated with daily
subcutaneous injection of C60-C3 or color-matched food coloring until death. Values are
means ± standard error of mean, *p < 0.05 by t test, n = 9, C60-C3 versus n = 7, vehicle. Adapted
from Ref. [6], Copyright 2004, with permission from Elsevier
12 2 Carbon-Based Nanomaterials for Nanozymes

2.1.3 Fullerene Derivatives as Peroxidase Mimics

A few studies also showed that fullerenes could mimic peroxidase [14, 15]. For
instance, the peroxidase-mimicking activity of C60[C(COOH)2]2 was demonstrated
by its catalyzed oxidation of 3,3′,5,5′-tetramethylbenzidine (TMB) with H2O2 [15].

2.2 Graphene and Derivatives

Peroxidase catalyzes the oxidation of its substrates with H2O2 into oxidized
products. The products are usually either colored or fluorescent, which enables
peroxidase for a wide range of bioanalytical and biomedical applications. The
peroxidase-mimicking activities of nanomaterials were initially studied with Fe3O4
nanoparticles by Yan and coworkers in 2007 [16]. Inspired by Yan’s work, Wang
et al. developed a general sensing strategy using Fe3O4 nanoparticles-based per-
oxidase mimic in 2008 (see detailed discussion in Chap. 4) [17]. Since then, lots of
different nanomaterials have been investigated to mimic peroxidase. Among them,
graphene and its various derivatives have showed great promise in mimicking
peroxidase.
Graphene and its derivatives with peroxidase-mimicking activities can be
roughly classified into two types. For the first type, the activities are solely from
graphene or its derivatives. For the second type, the activities are either from the
catalysts assembled onto graphene (or its derivatives) or from the synergistic effects
of both the decorated catalysts and graphene (or its derivatives). Note: other
enzyme-mimicking activities of graphene and its derivatives still remain to be
investigated [18].

2.2.1 Graphene and Its Derivatives as Peroxidase Mimics

The peroxidase-mimicking activities have been studied mainly with graphene


derivatives since pure graphene without any modifications is not water soluble. Qu
and coworkers reported the intrinsic peroxidase-mimicking activity of graphene
oxide with carboxyl modifications (GO-COOH) [19]. The activity was first
demonstrated by the catalytic oxidation of TMB with H2O2 in the presence of
GO-COOH (Fig. 2.4). The kinetic studies revealed that GO-COOH had higher
affinity toward TMB in comparison with natural peroxidase. Interestingly, the
GO-COOH catalyzed reactions proceeded via a ping-pong mechanism, which was
the same as that for natural peroxidase. Since they did not detect trace amount of
metal catalysts, they attributed the observed catalytic activity to the GO-COOH
itself. No mechanism responsible for the catalytic activity was proposed, though it
suggested that electron transfer from GC-COOH to H2O2 may be involved. By
2.2 Graphene and Derivatives 13

Fig. 2.4 Carboxyl-modified graphene oxide as peroxidase mimic. a Typical photographs of the
reaction solutions incubated at room temperature in pH 5.0 phosphate buffer (from left to right):
(1) 50 mM H2O2 and 800 μM TMB, colorless; (2) 40 mg/mL GO-COOH, black; (3) 50 mM
H2O2, 800 mM TMB and 40 mg mL/mL GO-COOH, turning blue. b The time-dependent
absorbance changes at 652 nm in the absence (black) or presence of different concentrations of
GO-COOH in pH 5.0 phosphate buffer at room temperature. Reprinted from Ref. [19], Copyright
2010, with permission from John Wiley and Sons

combining with natural glucose oxidase (GOx), the glucose in diluted blood and
fruit juice samples was successfully detected using the GC-COOH-based peroxi-
dase mimic [19].
Graphene derivatives and many other carbon-based nanomaterials have rich
oxygenated functional moieties (such as hydroxyl, ketone, carboxyl, epoxide, etc.).
These functional moieties may play key roles in their enzyme-mimicking activities.
To figure out the possible functional moieties responsible for the peroxidase-
mimicking activity of a graphene derivative called graphene quantum dots (GQDs),
several reagents were employed to selectively deactivate these functional moieties
(Fig. 2.5) [20]. GQDs are small pieces of graphene derivative with fluorescent
properties. It has ketone, hydroxyl, and carboxyl groups on the surface. These three
groups can selectively react with phenylhydrazine (PH), benzoic anhydride (BA),
and 2-bromo-1-phenylethanone (BrPE) respectively (Fig. 2.5). After treatment with
PH, BA, BrPE, the peroxidase-mimicking activities of the treated GQDs were
studied. As shown in Fig. 2.5b, the activity of PH-treated GQDs was significantly
inhibited while the activity of BA-treated GQDs was enhanced. The activity of
BrPE-treated GQDs remained almost the same as the untreated GQDs. Using a HO∙
specific fluorescent probe, the formation of HO∙ was confirmed during the catalytic
reaction. Further kinetic measurements and theoretical calculation were also carried
out. Taking together, these results suggested that ketone groups were the catalyti-
cally active sites and the carboxyl groups were the substrate binding sites. The
hydroxyl groups, on the other hand, played an inhibitory role. It may be applicable
to other carbon nanomaterials-based peroxidase mimics, which also have such
oxygenated moieties.
14 2 Carbon-Based Nanomaterials for Nanozymes

Fig. 2.5 Deciphering


peroxidase-mimicking
activity of GQDs. a Reactions
involved in selectively
deactivating functional
moieties on GQDs. b Relative
catalytic activities of GQDs
treated with different reagents.
Adapted from Ref. [20],
Copyright 2015, with
permission from John Wiley
and Sons

Since HO∙ was involved in peroxidase-mimicking activity of GQDs, they have


showed antibacterial activity even in the low level of H2O2. Both Gram-positive
(Staphylococcus aureus) and Gram-negative (Escherichia coli) bacteria could be
inhibited with the GQDs. The in vivo antibacterial efficacy was then evaluated
using a mouse injury model, showing that the combination of GQDs with H2O2
exhibited the best therapeutic effects compared with saline, H2O2, and GQDs [21].
The water solubility and stability of GO could be further enhanced by coating
with polymers. To this end, chitosan, a cationic polysaccharide, was used to coat
GO. The obtained chitosan–GO showed improved stability toward catalytic oxi-
dation of TMB with H2O2. Interestingly, it was found that the peroxidase-
mimicking activity of the chitosan–GO was regulated by light [22]. Under visible
light irradiation, the catalytic activity was turned on. More, the coated chitosan
could interact with concanavalin A (Con A) via a multivalent manner, and the
interaction would induce the aggregation of chitosan–GO. Such aggregation in turn
reduced the activity of the nanozyme. Glucose, however, could compete for the
2.2 Graphene and Derivatives 15

binding sites of chitosan on Con A due to the stronger interaction between glucose
and Con A. Therefore, the presence of glucose would disassemble the chitosan–
GO/Con A aggregates and thus recover the catalytic activity. Based on this phe-
nomenon, a facile phototriggered method for glucose detection was developed.
A linear range of 2.5–5.0 mM for glucose was achieved [22].

2.2.2 Decorated Graphene (or Its Derivatives) as Peroxidase


Mimics

The large surface area of graphene and its derivatives provides a good opportunity
to decorate them with various functional molecules and nanomaterials. As men-
tioned above, for such decorated graphene (or its derivatives), the peroxidase-
mimicking activities could be originated from either the assembled catalyst itself or
the catalyst/graphene (or its derivatives) assembles.
In their seminal report, Dong and coworkers assembled hemin onto reduced GO
(denoted as rGO) to form hemin–rGO complex and studied its peroxidase-
mimicking activity (Fig. 2.6). The rGO was obtained by reducing GO with
hydrazine. Then the hemin–rGO was prepared through the π-π stacking interaction
between hemin and rGO. Compared with hemin–rGO, rGO showed almost negli-
gible activity. Therefore, the peroxidase-mimicking activity of the hemin–rGO was
mainly attributed to the assembled hemin. As shown in Fig. 2.6, the hemin–
rGO-based nanozyme could catalyze the oxidation of TMB, ABTS (2,2′-azinobis
(3-ethylbenzothiozoline)-6-sulfonic acid, and OPD (o-phenylenediamine) into the
corresponding colored products with H2O2. Kinetic studies revealed that the
hemin–rGO catalyzed reaction also followed a ping-pong mechanism [23].

Fig. 2.6 Hemin-decorated rGO as peroxidase mimic. a Schematic illustration of peroxidase-like


activity of hemin–rGO. b Hemin–rGO catalyzed oxidation of various peroxidase substrates with
H2O2 to the corresponding colored products. Adapted from Ref. [23], Copyright 2011, with
permission from American Chemical Society
16 2 Carbon-Based Nanomaterials for Nanozymes

It is known that single-stranded DNA (ssDNA) and double-stranded DNA


(dsDNA) have different affinity toward various nanomaterials [24]. ssDNA binds
tightly onto graphene (or its derivatives) while dsDNA binds weakly. Such a dif-
ference could be further amplified by salt-induced aggregation. In the presence of
high concentration of salt (such as NaCl), ssDNA protects graphene (or its
derivatives) from aggregation while dsDNA cannot. Based on this phenomenon,
Dong et al. went on further to develop a label-free colorimetric method for
single-nucleotide polymorphisms (SNPs) (Fig. 2.7) [23]. The probe ssDNA could
stabilize the hemin–rGO and thus retained its peroxidase-mimicking activity. When
the complementary target ssDNA was hybridized with the probe ssDNA, the
formed dsDNA could not stabilize the hemin–rGO, which led to significant inhi-
bition of its peroxidase-mimicking activity and produced the weakest signal. When
a target ssDNA with a single-base mismatch was introduced, the formed duplex
could partially protect the hemin–rGO from aggregation and thus retained part of its
peroxidase-mimicking activity. This in turn resulted in the signal with medium
intensity. As shown in Fig. 2.7, one could easily distinguish the single-base mis-
matched target DNA from the complementary one even with naked eyes [23].
This sensing strategy could be applicable to functional nucleic acids (such as
aptamers) [24, 25]. An aptamer is an ssDNA or ssRNA that can specifically bind to
its target. The binding usually induces a conformational change of the aptamer.
Such a conformational change could in principle be sensed with hemin–rGO. For

Fig. 2.7 Hemin-decorated rGO as peroxidase mimic for single-nucleotide polymorphisms.


a Protocol for SNPs detection. (a) Probe ssDNA (no precipitation, dark blue), (b) single-base
mismatched duplex DNA (small amount of precipitation, blue), and (c) complementary duplex
DNA (much precipitation, light blue). b Time-dependent absorbance changes in the presence of
different amounts of target ssDNA. c Time-dependent absorbance changes with corresponding
supernatant in (a) ssDNA, (b, c, d) single-base mismatched duplex DNA, and (e) complementary
duplex DNA. Reprinted from Ref. [23], Copyright 2011, with permission from American
Chemical Society
2.2 Graphene and Derivatives 17

instance, when an aptamer for acetamiprid (an insecticide) was used to stabilize
hemin–rGO, the nanozyme showed high activity. However, the presence of acet-
amiprid would form the acetamiprid-aptamer complex, which did not protect the
hemin–rGO efficiently. Therefore, the nanozyme’s activity was significantly
inhibited due to the aggregation. With this sensing strategy, as low as 40 nM
acetamiprid was detected [26]. Other targets of interests can also be detected when
the corresponding aptamers are used.
The synergistic effects were observed for numerous decorated graphene (or
its derivatives) [27–35]. Among GO, rGO, AuNPs, the mixture of rGO and
AuNPs, and the gold nanoparticles-decorated rGO (denoted as AuNPs@rGO),
AuNPs@rGO exhibited the highest peroxidase-mimicking activity. As shown in
Fig. 2.8, the significantly enhanced catalytic activity of the AuNPs@rGO was
attributed to the synergistic effects. Careful study suggested that the synergistic
effects were due to: (a) the strong interaction between Au 5d of AuNPs and C 2p of

Fig. 2.8 AuNPs@rGO with synergistic peroxidase-mimicking activity and its use for DNA
sensing. a TEM (transmission electron microscopy) image of AuNPs@rGO.
b Peroxidase-mimicking activities of various nanomaterials. c Comparison of fluorescent retention
ratio of FAM-labeled probe ssDNA and the corresponding dsDNA after incubated with
AuNPs@rGO. d Time-dependent absorbance changes with varying concentrations of target
ssDNA. Adapted from Ref. [27], Copyright 2012, with permission from American Chemical
Society
18 2 Carbon-Based Nanomaterials for Nanozymes

rGO at the interface, which was favorable to H2O2 and HO∙ absorption; and (b) the
modified electronic structure and Fermi level of rGO owing to the above-mentioned
interfacing, which in turn resulted in the n-type doping of rGO and accelerated the
catalytic reactions [27]. More, by further exploring the different affinities of ssDNA
and dsDNA toward the AuNPs@rGO, a general analytical strategy for target DNA
was developed (Fig. 2.8). The ssDNA specific nuclease (such as S1 nuclease) was
also successfully detected using the developed method, where the presence of S1
nuclease would cleave the probe ssDNA into small fragments and thus recover the
nanozyme’s activity. If an aptamer was used as the probe ssDNA, the corre-
sponding target (such as insulin) could be detected [27].
A hybrid called GSF@AuNPs with peroxidase-mimicking activity was prepared
by in situ formation of AuNPs onto sandwich-like mesoporous silica/rGO. The
silica/rGO was pre-conjugated with folic acid for tumor cell recognition [29]. The
nanozyme was then used for detection of cancer cells with overexpressed folate
receptor. For instance, rapid detection of HeLa cells was achieved. More, since HO∙
was generated during the catalysis, the nanozyme was also used to selectively kill
cancer cells with the help of either exogenous or endogenous H2O2 [29].
Chen and coworkers prepared PtNPs decorated GO (i.e., PtNPs/GO) and further
modified it with folic acid [36]. Based on the peroxidase-mimicking activity of the
folic acid-PtNPs/GO, a colorimetric assay for cancer cells was developed (Fig. 2.9).
With the developed assay, as few as 125 cancer cells have been detected by naked
eyes.
By conjugating the monoclonal antibody for aflatoxin B1 with a
peroxidase-mimicking nanozyme, an electrochemical immunoassay for aflatoxin B1
was reported by Tang et al. [32]. The nanozyme had a structure of
PtNPs/CoTPP/rGO, in which CoTPP was 5,10,15,20-tetraphenyl-21H,23H-
porphine cobalt. Aflatoxin B1 is a highly toxic metabolite secreted by food fungi
and its detection still needs rapid methods with high sensitivity and low cost. With

Fig. 2.9 Colorimetric detection of cancer cells with folic acid–PtNPs/GO as a peroxidase mimic.
Reprinted from Ref. [36], Copyright 2014, with permission from American Chemical Society
2.2 Graphene and Derivatives 19

the developed immunoassay, as low as 5.0 pg/mL of aflatoxin B1 was successfully


detected. Besides, the immunoassay has been used for analyzing real samples, such
as the naturally contaminated peanut samples, showing good agreement with
ELISA (enzyme-linked immunosorbent assay) kit [32].
Yu and coworkers reported a disposable electrochemical immunosensor based
on peroxidase-mimicking nanozyme for cancer antigen 153 (CA153) detection
[34]. Their nanozyme had a structure of ZnFe2O4@silica/GO (Fig. 2.10). The
antibodies were conjugated onto the silica shell. Using a sandwich assay format,
sensitive and selective detection of CA153 was achieved with a dynamic range
from 10−3 to 200 U/mL and a detection limit of 2.8 × 10−4 U/mL. The fabricated
immunosensor was further used to detect CA153 in serum samples and the results
were consistent with the clinical ones.

Fig. 2.10 Peroxidase-mimicking nanozyme and its use for disposable electrochemical
immunosensor. Reprinted from Ref. [34], Copyright 2014, with permission from Elsevier
20 2 Carbon-Based Nanomaterials for Nanozymes

2.3 Carbon Nanotubes

CNTs as well as decorated CNTs have been used to mainly mimic peroxidase
though other CNT-based enzyme mimics were also reported [37, 38].

2.3.1 Carbon Nanotubes as Peroxidase Mimics

Qu and coworkers reported the peroxidase-mimicking activity of single-walled


carbon nanotubes (SWNTs) [39]. The activity of SWNTs was investigated by
catalytic oxidation of TMB with H2O2. Since metal catalysts are usually used to
grow SWNTs, a trace amount of metal catalyst residues rather than SWNTs
themselves may be responsible for the mimicking activity. To address this concern,
the sonication-assisted washing with mixed acids (i.e., a mixture of concentrated
sulfuric and nitric acids) was carried out to completely remove the metal residues
(i.e., Co). It was found that pristine SWNTs and treated SWNTs did not show any
significant differences in their catalytic activities. This confirmed that the
peroxidase-mimicking activity of SWNTs was from the SWNTs themselves instead
of the metal residues. By exploring the different affinities of ssDNA and dsDNA
toward SWNTs, they developed a colorimetric assay for DNA detection [39].
It should be noted that in the above study, only the effect of Co on SWNTs’
peroxidase-mimicking activity was tested. Other metal residues may have different
effects. Zhu and coworkers indeed found that Fe content in the helical CNTs played
an important role in their catalytic activities [40]. As shown in Fig. 2.11, the more
Fe content of helical CNT was, the higher its peroxidase-mimicking activity of
helical CNT was. Even for the helical CNT with the lowest amount of Fe, its
activity was still higher than that of MWNTs (multiwalled CNTs). Zhu’s and Qu’s
results suggest that more systematic studies are needed to decipher the exact
mechanisms of CNTs’ enzyme-mimicking activities. Zhu et al. then fabricated an
electrochemical sensor using the helical CNTs as peroxidase mimic for H2O2
detection [40].
Like graphene, the decoration of CNTs could also synergistically enhance their
peroxidase-mimicking activities. When MWNTs were decorated with magnetic
silica nanoparticles, the decorated MWNTs exhibited higher peroxidase-mimicking
activity compared with the individual components (i.e., MWNTs and magnetic
silica nanoparticles, respectively) [41]. Since the decoration of magnetic silica
nanoparticles onto MWNTs was achieved by the Cu2+-mediated click chemistry, a
colorimetric assay for Cu2+ was proposed (Fig. 2.12). The decorated MWNTs
could be concentrated by a magnet and then used to oxidize TMB to its colored
products. On the other hand, the undecorated MWNTs would be washed away and
only the magnetic silica nanoparticles would be concentrated by a magnet. The
latter exhibited much lower catalytic activity compared with the decorated
2.3 Carbon Nanotubes 21

Fig. 2.11 Peroxidase-mimicking activity of helical CNTs. a SEM (scanning electron microscopy)
image of helical CNTs. b Peroxidase-mimicking activities of helical CNTs with different Fe
contents as well as MWNTs. The catalytic reaction without CNTs was shown as a control.
Adapted from Ref. [40], Copyright 2011, with permission from John Wiley and Sons

MWNTs. Via such a sensing strategy, high sensitive and selective detection of Cu2+
has been carried out [41].
Other decorated CNTs with synergistically enhanced peroxidase-mimicking
activities have also been developed and used to detect biologically important targets

Fig. 2.12 Cu2+ detection using magnetic silica nanoparticles clicked on MWNTs. a The sensing
mechanism. b The time-dependent absorbance changes in the absence (black) or presence of
different concentration of Cu2+. c Calibration curve for variable concentrations of Cu2+. The error
bars represent the standard deviation of three measurements. Reprinted from Ref. [41], Copyright
2010, with permission from Royal Society of Chemistry
22 2 Carbon-Based Nanomaterials for Nanozymes

(also see Tables A1 and A2). For example, when MWNTs were filled with Prussian
blue nanoparticles, the formed nanozyme has been used for colorimetric detection
of H2O2 [42]. When the nanozyme was further combined with glucose oxidase, it
has been successfully used for glucose detection with a linear range of 1 μM to
1.0 mM and a detection limit of 200 nM. The potential practical application was
demonstrated by detecting glucose in serum samples, showing satisfactory recov-
eries of 94–106 % [42]. Cholesterol levels in milk powder were evaluated by using
ZnO nanoparticles-decorated CNTs as a peroxidase mimic [43]. Cholesterol was
catalytically oxidized with cholesterol oxidase to produce H2O2, which was then
used for oxidizing ABTS into the colored product with the nanozyme.
A colorimetric approach to detection of D-alanine was developed by using Au
nanoparticles-decorated SWNTs as a peroxidase mimic [44]. D-alanine was cat-
alytically oxidized with D-amino acids oxidase to produce H2O2 for the further
oxidation of TMB. The proposed approach showed high selectivity and high sen-
sitivity toward D-alanine detection.
A paper-based immunoassay was developed using ZnFe2O4-decorated MWNTs
as a peroxidase mimic [45]. Carcinoembryonic antigen (CEA) was chosen due to its
correlation with cancer. The nanozymes-based immunoassay exhibited high sen-
sitivity and robustness. More, the immunoassay has been used to detect CEA in
clinical serum samples, showing consistent results compared with a commercialized
ELISA method [45].

2.3.2 Carbon Nanotubes as Other Enzyme Mimics

As discussed above, it has been established that fullerenes can efficiently scavenge
radicals due to their SOD-mimicking activities. Thus, it is reasonable to investigate
the radical scavenging activities of CNTs. Tour and coworkers have studied the
radical scavenging activities of several SWNTs (Fig. 2.13) [37]. They used butylated
hydroxytoluene (BHT), a phenolic antioxidant, to modify the SWNTs. The anchored
BHT would endow the SWNTs with radical scavenging activities. Pristine SWNT
(SWNT-1) was not water soluble. Therefore, it was solubilized by either wrapping a
polymer (such as Pluronic for SWNT-2) or introducing carboxyl moieties by mixed
acid treatment/cleavage (i.e., SWNT-3). SWNT-3 was further modified with PEG
(poly(ethylene glycol)) to produce SWNT-4, which was soluble in buffer.
The radical scavenging activities were evaluated by comparing with Trolox, a
vitamin E derivate. The TME (Trolox mass equivalence) values were then deter-
mined. Unexpectedly, SWNT-4 without BHT modification exhibited nearly 40
times higher activity (Fig. 2.14). This indicated that SWNTs alone could act as
antioxidants. For SWNT-5, amine-BHT was assembled via electrostatic interac-
tions while amine-BHT was covalently conjugated for SWNT-6. Since electrostatic
binding was more efficient than the covalent binding, more amine-BHT should bind
onto SWNT-5 than SWNT-6. As expected, SWNT-5 showed higher radical
scavenging activity than SWNT-6 did (Fig. 2.14). For SWNT-2 and SWNT-7, one
2.3 Carbon Nanotubes 23

Fig. 2.13 SWNTs used for scavenging radicals. Adapted from Ref. [37], Copyright 2009, with
permission from American Chemical Society

would expect that SWNT-7 would have higher activity since it had extra BHT
moieties. Surprisingly, SWNT-7 exhibited lower activity when compared with
SWNT-2 (Fig. 2.14) [37]. This unexpected result indicated that pristine SWNTs
had higher radical scavenging activities than BHT. This study demonstrated that
SWNTs could act as potent antioxidant without acute cellular toxicity.

Fig. 2.14 Radical scavenging activities (i.e., the TME values) of the SWNTs studied. C60-5 was
also included for a comparison. Adapted from Ref. [37], Copyright 2009, with permission from
American Chemical Society
24 2 Carbon-Based Nanomaterials for Nanozymes

Note that though it has established that fullerenes act as SOD mimic to scavenge
radicals, whether CNTs share a similar SOD mimetic mechanism still remains to be
investigated.

2.4 Other Carbon-Based Nanomaterials

Other carbon-based nanomaterials (such as carbon nanohorn, carbon nanodots,


carbon nanoclusters, etc.) have also been used to mimic peroxidase and other
enzymes [46–58].

2.4.1 Other Carbon Nanomaterials as Peroxidase Mimics

Several groups studied the peroxidase-mimicking activities of carbon nanohorns,


carbon nanodots, etc. [46–48, 50–52, 58, 59]
For instance, Xu and coworkers demonstrated that carboxyl-functionalized
single-walled carbon nanohorns exhibited peroxidase-like activity (Fig. 2.15) [46].
When the nanozyme was further combined with glucose oxidase, a facile colori-
metric assay for glucose was developed. Carbon nanodots with an average size of
2.5 nm also showed peroxidase-mimicking activity and have been used for glucose
sensing (Fig. 2.15c) [51]. The peroxidase-like activity of selenium-doped graphitic
carbon nitride nanosheets was demonstrated and further explored for xanthine
detection when facilitated with xanthine oxidase [52].
Using [Cu3(BTC)2] (BTC = 1,3,5-benzene tricarboxylate) as a precursor, copper
nanoparticles-decorated carbon nanocomposite was fabricated via a one-pot ther-
molysis method (Fig. 2.15d) [47]. It demonstrated that the nanocomposite exhibited
peroxidase mimetic activity. Ascorbic acid, a biologically important antioxidant,
could competitively inhibit the catalytic oxidation of peroxidase substrate (such as
TMB in this study). Based on this inhibition phenomenon, a colorimetric method
for ascorbic acid was developed [47, 60]. The ascorbic acid content in tablets has
been successfully determined with the nanozyme-based method [47].

2.4.2 Other Carbon Nanomaterials as SOD Mimics

The SOD-mimicking activities of nitrogen-doped carbon nanodots have been


studied [58]. It demonstrated that primary amine were the optimal reagents for
nitrogen doping. When the obtained nitrogen-doped carbon nanodots were added to
H2O2-treated cells, they would enhance the cell viability in a concentration-
dependent manner. It suggested that the nitrogen-doped carbon nanotdots protect
cells from H2O2-induced injury by eliminating the ROS and stimulating the native
SOD expression [58].
References 25

Fig. 2.15 Carbon nanomaterials as peroxidase mimics. a Single-walled carbon nanohorns as


peroxidase mimic and b their use for glucose detection. c Carbon nanodots as peroxidase mimics.
d Copper nanoparticles-decorated carbon as peroxidase mimics. a and b Reprinted from Ref. [46],
Copyright 2015, with permission from Royal Society of Chemistry. c Reprinted from Ref. [51],
Copyright 2011, with permission from Royal Society of Chemistry. d Reprinted from Ref. [47],
Copyright 2014, with permission from John Wiley and Sons

References

1. Tokuyama, H., Yamago, S., Nakamura, E., Shiraki, T., & Sugiura, Y. (1993). Photoinduced
biochemical-activity of fullerene carboxylic-acid. Journal of the American Chemical Society,
115, 7918–7919.
2. Nakamura, E., & Isobe, H. (2003). Functionalized fullerenes in water. The first 10 years of
their chemistry, biology, and nanoscience. Accounts of Chemical Research, 36, 807–815.
3. Boutorine, A. S., Tokuyama, H., Takasugi, M., Isobe, H., Nakamura, E., & Helene, C. (1994).
Fullerene-oligonucleotide conjugates: Photoinduced sequence-specific DNA cleavage.
Angewandte Chemie-International Edition, 33, 2462–2465.
4. Yamakoshi, Y. N., Yagami, T., Sueyoshi, S., & Miyata, N. (1996). Acridine adduct of 60
fullerene with enhanced DNA-cleaving activity. Journal of Organic Chemistry, 61,
7236–7237.
5. Dugan, L. L., Gabrielsen, J. K., Yu, S. P., Lin, T. S., & Choi, D. W. (1996).
Buckminsterfullerenol free radical scavengers reduce excitotoxic and apoptotic death of
cultured cortical neurons. Neurobiology of Disease, 3, 129–135.
6. Ali, S. S., Hardt, J. I., Quick, K. L., Kim-Han, J. S., Erlanger, B. F., Huang, T. T., et al. (2004).
A biologically effective fullerene (C-60) derivative with superoxide dismutase mimetic
properties. Free Radical Biology and Medicine, 37, 1191–1202.
26 2 Carbon-Based Nanomaterials for Nanozymes

7. Krusic, P. J., Wasserman, E., Keizer, P. N., Morton, J. R., & Preston, K. F. (1991). Radical
reactions of C60. Science, 254, 1183–1185.
8. Dugan, L. L., Turetsky, D. M., Du, C., Lobner, D., Wheeler, M., Almli, C. R., et al. (1997).
Carboxyfullerenes as neuroprotective agents. Proceedings of the National Academy of
Sciences of the United States of America, 94, 9434–9439.
9. Liu, G.-F., Filipovic, M., Ivanovic-Burmazovic, I., Beuerle, F., Witte, P., & Hirsch, A. (2008).
High catalytic activity of dendritic C60 monoadducts in metal-free superoxide dismutation.
Angewandte Chemie-International Edition, 47, 3991–3994.
10. Osuna, S., Swart, M., & Sola, M. (2010). On the mechanism of action of fullerene derivatives
in superoxide dismutation. Chemistry-A European Journal, 16, 3207–3214.
11. Chen, T., Li, Y.-Y., Zhang, J.-L., Xu, B., Lin, Y., Wang, C.-X., et al. (2011). Protective effect
of C60-methionine derivate on lead-exposed human SH-SY5Y neuroblastoma cells. Journal of
Applied Toxicology, 31, 255–261.
12. Gharbi, N., Pressac, M., Hadchouel, M., Szwarc, H., Wilson, S. R., & Moussa, F. (2005).
60 Fullerene is a powerful antioxidant in vivo with no acute or subacute toxicity. Nano Letters,
5, 2578–2585.
13. Baati, T., Bourasset, F., Gharbi, N., Njim, L., Abderrabba, M., Kerkeni, A., et al. (2012). The
prolongation of the lifespan of rats by repeated oral administration of 60 fullerene.
Biomaterials, 33, 4936–4946.
14. Okuda, K., Mashino, T., & Hirobe, M. (1996). Superoxide radical quenching and cytochrome
c peroxidase-like activity of C60-dimalonic acid, C62(COOH)4. Bioorganic & Medicinal
Chemistry Letters, 6, 539–542.
15. Li, R. M., Zhen, M. M., Guan, M. R., Chen, D. Q., Zhang, G. Q., Ge, J. C., et al. (2013).
A novel glucose colorimetric sensor based on intrinsic peroxidase-like activity of C60-
carboxyfullerenes. Biosensors & Bioelectronics, 47, 502–507.
16. Gao, L. Z., Zhuang, J., Nie, L., Zhang, J. B., Zhang, Y., Gu, N., et al. (2007). Intrinsic
peroxidase-like activity of ferromagnetic nanoparticles. Nature Nanotechnology, 2, 577–583.
17. Wei, H., & Wang, E. (2008). Fe3O4 magnetic nanoparticles as peroxidase mimetics and their
applications in H2O2 and glucose detection. Analytical Chemistry, 80, 2250–2254.
18. Liu, T., Niu, X., Shi, L., Zhu, X., Zhao, H., & Lana, M. (2015). Electrocatalytic analysis of
superoxide anion radical using nitrogen-doped graphene supported Prussian Blue as a
biomimetic superoxide dismutase. Electrochimica Acta, 176, 1280–1287.
19. Song, Y., Qu, K., Zhao, C., Ren, J., & Qu, X. (2010). Graphene oxide: Intrinsic peroxidase
catalytic activity and its application to glucose detection. Advanced Materials, 22, 2206–2210.
20. Sun, H. J., Zhao, A. D., Gao, N., Li, K., Ren, J. S., & Qu, X. G. (2015). Deciphering a
nanocarbon-based artificial peroxidase: Chemical identification of the catalytically active and
substrate-binding sites on graphene quantum dots. Angewandte Chemie-International Edition,
54, 7176–7180.
21. Sun, H., Gao, N., Dong, K., Ren, J., & Qu, X. (2014). Graphene quantum dots-band-aids used
for wound disinfection. ACS Nano, 8, 6202–6210.
22. Wang, G. L., Xu, X. F., Wu, X. M., Cao, G. X., Dong, Y. M., & Li, Z. J. (2014).
Visible-light-stimulated enzymelike activity of graphene oxide and its application for facile
glucose sensing. Journal of Physical Chemistry C, 118, 28109–28117.
23. Guo, Y. J., Deng, L., Li, J., Guo, S. J., Wang, E. K., & Dong, S. J. (2011). Hemin-graphene
hybrid nanosheets with intrinsic peroxidase-like activity for label-free colorimetric detection of
single-nucleotide polymorphism. ACS Nano, 5, 1282–1290.
24. Wei, H., Li, B. L., Li, J., Wang, E. K., Dong, S. J. (2007) Simple and sensitive aptamer-based
colorimetric sensing of protein using unmodified gold nanoparticle probes. Chemical
Communications, 3735–3737.
25. Wei, H., Li, B. L., Li, J., Dong, S. J., & Wang, E. K. (2008). DNAzyme-based colorimetric
sensing of lead (Pb2+) using unmodified gold nanoparticle probes. Nanotechnology, 19,
095501.
References 27

26. Yang, Z., Qian, J., Yang, X., Jiang, D., Du, X., Wang, K., et al. (2015). A facile label-free
colorimetric aptasensor for acetamiprid based on the peroxidase-like activity of
hemin-functionalized reduced graphene oxide. Biosensors & Bioelectronics, 65, 39–46.
27. Liu, M., Zhao, H. M., Chen, S., Yu, H. T., & Quan, X. (2012). Interface engineering catalytic
graphene for smart colorimetric biosensing. ACS Nano, 6, 3142–3151.
28. Liu, M., Zhao, H. M., Chen, S., Yu, H. T., & Quan, X. (2012). Stimuli-responsive peroxidase
mimicking at a smart graphene interface. Chemical Communications, 48, 7055–7057.
29. Maji, S. K., Mandal, A. K., Nguyen, K. T., Borah, P., & Zhao, Y. L. (2015). Cancer cell
detection and therapeutics using peroxidase-active nanohybrid of gold nanoparticle-loaded
mesoporous silica-coated graphene. ACS Applied Materials & Interfaces, 7, 9807–9816.
30. Wang, H., Li, S., Si, Y. M., Sun, Z. Z., Li, S. Y., & Lin, Y. H. (2014). Recyclable enzyme
mimic of cubic Fe3O4 nanoparticles loaded on graphene oxide-dispersed carbon nanotubes
with enhanced peroxidase-like catalysis and electrocatalysis. Journal of Materials
Chemistry B, 2, 4442–4448.
31. Dutta, S., Ray, C., Mallick, S., Sarkar, S., Sahoo, R., Negishi, Y., et al. (2015). A gel-based
approach to design hierarchical cus decorated reduced graphene oxide nanosheets for
enhanced peroxidase-like activity leading to colorimetric detection of dopamine. Journal of
Physical Chemistry C, 119, 23790–23800.
32. Shu, J., Qiu, Z., Wei, Q., Zhuang, J., & Tang, D. (2015). Cobalt-porphyrin-
platinum-functionalized reduced graphene oxide hybrid nanostructures: A novel peroxidase
mimetic system for improved electrochemical immunoassay. Scientific Reports, 5, 15113.
33. Ma, Z., Qiu, Y. F., Yang, H. H., Huang, Y. M., Liu, J. J., Lu, Y., et al. (2015). Effective
synergistic effect of dipeptide-polyoxometalate-graphene oxide ternary hybrid materials on
peroxidase-like mimics with enhanced performance. ACS Applied Materials & Interfaces, 7,
22036–22045.
34. Ge, S., Sun, M., Liu, W., Li, S., Wang, X., Chu, C., et al. (2014). Disposable electrochemical
immunosensor based on peroxidase-like magnetic silica-graphene oxide composites for
detection of cancer antigen 153. Sensors and Actuators B-Chemical, 192, 317–326.
35. Yan, X., Gu, Y., Li, C., Tang, L., Zheng, B., Li, Y., et al. (2016). Synergetic catalysis based on
the proline tailed metalloporphyrin with graphene sheet as efficient mimetic enzyme for
ultrasensitive electrochemical detection of dopamine. Biosensors & Bioelectronics, 77,
1032–1038.
36. Zhang, L. N., Deng, H. H., Lin, F. L., Xu, X. W., Weng, S. H., Liu, A. L., et al. (2014). In situ
growth of porous platinum nanoparticles on graphene oxide for colorimetric detection of
cancer cells. Analytical Chemistry, 86, 2711–2718.
37. Lucente-Schultz, R. M., Moore, V. C., Leonard, A. D., Price, B. K., Kosynkin, D. V., Lu, M.,
et al. (2009). Antioxidant single-walled carbon nanotubes. Journal of the American Chemical
Society, 131, 3934–3941.
38. Zhang, B., He, Y., Liu, B., & Tang, D. (2014). NiCoBP-doped carbon nanotube hybrid: A
novel oxidase mimetic system for highly efficient electrochemical immunoassay. Analytica
Chimica Acta, 851, 49–56.
39. Song, Y., Wang, X., Zhao, C., Qu, K., Ren, J., & Qu, X. (2010). Label-free colorimetric
detection of single nucleotide polymorphism by using single-walled carbon nanotube intrinsic
peroxidase-like activity. Chemistry-A European Journal, 16, 3617–3621.
40. Cui, R., Han, Z., & Zhu, J.-J. (2011). Helical carbon nanotubes: Intrinsic peroxidase catalytic
activity and its application for biocatalysis and biosensing. Chemistry-A European Journal,
17, 9377–9384.
41. Song, Y., Qu, K., Xu, C., Ren, J., & Qu, X. (2010). Visual and quantitative detection of copper
ions using magnetic silica nanoparticles clicked on multiwalled carbon nanotubes. Chemical
Communications, 46, 6572–6574.
28 2 Carbon-Based Nanomaterials for Nanozymes

42. Wang, T., Fu, Y. C., Chai, L. Y., Chao, L., Bu, L. J., Meng, Y., et al. (2014). Filling carbon
nanotubes with prussian blue nanoparticles of high peroxidase- like catalytic activity for
colorimetric chemo and biosensing. Chemistry-A European Journal, 20, 2623–2630.
43. Hayat, A., Haider, W., Raza, Y., & Marty, J. L. (2015). Colorimetric cholesterol sensor based
on peroxidase like activity of zinc oxide nanoparticles incorporated carbon nanotubes.
Talanta, 143, 157–161.
44. Haider, W., Hayat, A., Raza, Y., Chaudhry, A. A., Ihtesham Ur, R., & Marty, J. L. (2015).
Gold nanoparticle decorated single walled carbon nanotube nanocomposite with synergistic
peroxidase like activity for D-alanine detection. RSC Advances, 5, 24853–24858.
45. Liu, W. Y., Yang, H. M., Ding, Y. A., Ge, S. G., Yu, J. H., Yan, M., et al. (2014). Paper-based
colorimetric immunosensor for visual detection of carcinoembryonic antigen based on the high
peroxidase-like catalytic performance of ZnFe2O4-multiwalled carbon nanotubes. Analyst,
139, 251–258.
46. Zhu, S. Y., Zhao, X. E., You, J. M., Xu, G. B., & Wang, H. (2015).
Carboxylic-group-functionalized single-walled carbon nanohorns as peroxidase mimetics
and their application to glucose detection. Analyst, 140, 6398–6403.
47. Tan, H. L., Ma, C. J., Gao, L., Li, Q., Song, Y. H., Xu, F. G., et al. (2014). Metal-organic
framework-derived copper nanoparticle@carbon nanocomposites as peroxidase mimics for
colorimetric sensing of ascorbic acid. Chemistry-a European Journal, 20, 16377–16383.
48. Liu, S., Tian, J. Q., Wang, L., Luo, Y. L., & Sun, X. P. (2012). A general strategy for the
production of photoluminescent carbon nitride dots from organic amines and their application
as novel peroxidase-like catalysts for colorimetric detection of H2O2 and glucose. RSC
Advances, 2, 411–413.
49. Wang, X. H., Qu, K. G., Xu, B. L., Ren, J. S., & Qu, X. G. (2011). Multicolor luminescent
carbon nanoparticles: Synthesis, supramolecular assembly with porphyrin, intrinsic
peroxidase-like catalytic activity and applications. Nano Research, 4, 908–920.
50. Dong, Y. M., Zhang, J. J., Jiang, P. P., Wang, G. L., Wu, X. M., Zhao, H., et al. (2015).
Superior peroxidase mimetic activity of carbon dots-Pt nanocomposites relies on synergistic
effects. New Journal of Chemistry, 39, 4141–4146.
51. Shi, W. B., Wang, Q. L., Long, Y. J., Cheng, Z. L., Chen, S. H., Zheng, H. Z., et al. (2011).
Carbon nanodots as peroxidase mimetics and their applications to glucose detection. Chemical
Communications, 47, 6695–6697.
52. Qian, F. M., Wang, J. M., Ai, S. Y., & Li, L. F. (2015). As a new peroxidase mimetics: The
synthesis of selenium doped graphitic carbon nitride nanosheets and applications on
colorimetric detection of H2O2 and xanthine. Sensors and Actuators B-Chemical, 216,
418–427.
53. Samuel, E. L. G., Marcano, D. C., Berka, V., Bitner, B. R., Wu, G., Potter, A., et al. (2015).
Highly efficient conversion of superoxide to oxygen using hydrophilic carbon clusters.
Proceedings of the National Academy of Sciences of the United States of America, 112,
2343–2348.
54. Bitner, B. R., Marcano, D. C., Berlin, J. M., Fabian, R. H., Cherian, L., Culver, J. C., et al.
(2012). Antioxidant carbon particles improve cerebrovascular dysfunction following traumatic
brain injury. ACS Nano, 6, 8007–8014.
55. Marcano, D. C., Bitner, B. R., Berlin, J. M., Jarjour, J., Lee, J. M., Jacob, A., et al. (2013).
Design of poly(ethylene glycol)-functionalized hydrophilic carbon clusters for targeted
therapy of cerebrovascular dysfunction in mild traumatic brain injury. Journal of
Neurotrauma, 30, 789–796.
56. Samuel, E. L. G., Duong, M. T., Bitner, B. R., Marcano, D. C., Tour, J. M., & Kent, T. A.
(2014). Hydrophilic carbon clusters as therapeutic, high-capacity antioxidants. Trends in
Biotechnology, 32, 501–505.
57. Nilewski, L. G., Sikkema, W. K. A., Kent, T. A., & Tour, J. M. (2015). Carbon nanoparticles
and oxidative stress: Could an injection stop brain damage in minutes? Nanomedicine, 10,
1677–1679.
References 29

58. Xu, Z.-Q., Lan, J.-Y., Jin, J.-C., Dong, P., Jiang, F.-L., & Liu, Y. (2015). Highly
photoluminescent nitrogen-doped carbon nanodots and their protective effects against
oxidative stress on cells. ACS Applied Materials & Interfaces, 7, 28346–28352.
59. Clark, A., Zhu, A. P., Sun, K., & Petty, H. R. (2011). Cerium oxide and platinum
nanoparticles protect cells from oxidant-mediated apoptosis. Journal of Nanoparticle
Research, 13, 5547–5555.
60. Cheng, H., Wang, X., & Wei, H. (2015). Ratiometric electrochemical sensor for effective and
reliable detection of ascorbic acid in living brains. Analytical Chemistry, 87, 8889–8895.
Chapter 3
Metal-Based Nanomaterials for Nanozymes

Abstract The use of metal nanomaterials for mimicking natural enzymes is dis-
cussed in this chapter. These nanozymes are roughly classified into two types: for
type I, the nanozymes’ activities are entirely from the assembled monolayer onto a
metallic core rather than the core itself; for type II, the nanozymes’ activities are
originated from the metal nanomaterials themselves. For both of them, their enzyme
mimetic activities (such as RNase mimics, DNase mimics, superoxide dismutase
mimics, peroxidase mimics, catalase mimics, etc.) are discussed. The catalytic
mechanisms for the multiple enzyme mimicking activities of metal nanomaterials
are elucidated by combing computational studies with experimental results.
Representative examples for applications, from biosensing and immunoassays to
bioimaging and therapeutics, are covered.

 
Keywords Nanozymes Artificial enzymes Enzyme mimics Metal nanoma- 
 
terials Multiple enzyme mimics Catalytic monolayer-protected gold nanopar-
  
ticles Immunoassays Self-assembly Computational study Bioanalysis 

The metal-based nanozymes could be roughly classified into two types: for type I,
the nanozymes’ activities are entirely from the assembled monolayer onto a metallic
core rather than the core itself; for type II, the nanozymes’ activities are originated
from the metal nanomaterials themselves. In this chapter, both of them are dis-
cussed to highlight their various enzyme mimicking activities and wide
applications.

3.1 Metal Nanomaterials with Catalytic


Monolayers (Type I)

Metal nanomaterials (such as Au, Ag, etc.) with self-assembled monolayers (par-
ticularly the thiolated monolayers) have been extensively explored due to their great
importance to nanotechnology [1]. If catalytic moieties were introduced into the

© The Author(s) 2016 31


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_3
32 3 Metal-Based Nanomaterials for Nanozymes

monolayers, one would expect the metal nanomaterials protected by the monolayers
to be catalytically active. These functionalized metal nanoparticles could indeed
exert enzyme-like catalysis, and therefore, they have been regarded as nanozymes
[2–5]. These nanozymes could be further classified into three subtypes according to
the molecules used for the monolayers: the first one uses alkanethiol terminated
with catalytic moieties; the second one uses alkanethiol without catalytic terminus,
and the catalytic moieties, instead, are further assembled onto (or into) the mono-
layers; and the third one uses thiolated catalytic biomolecules (such as thiolated
DNAzymes).

3.1.1 AuNPs Protected by Alkanethiol with


Catalytic Terminal Moieties

(a) Alkanethiol-protected AuNPs as RNase mimics: activities and mechanisms


As mentioned in Chap. 1, in their initial study, Scrimin and co-workers used
alkanethiol terminated with catalytic moieties to modify the AuNP core (Fig. 3.1).
The obtained AuNPs assemblies were then used as a metallonuclease (i.e., RNase)
mimic to catalyze the transphosphorylation of 2-hydroxypropyl p-nitrophenyl
phosphate (HPNPP) [2].

Fig. 3.1 Alkanethiol-protected AuNPs as metallonuclease mimics. a Transphosphorylation


catalyzed by RNase A. b Structure of AuNP-1 nanozyme. c Transphosphorylation of HPNPP
catalyzed by AuNP-1 nanozyme. c Adapted from Ref. [2], Copyright 2004, with permission from
John Wiley and Sons
3.1 Metal Nanomaterials with Catalytic Monolayers (Type I) 33

As shown in Fig. 3.2, in comparison with the uncatalyzed transphosphorylation


of HPNPP, the reaction has been accelerated by more than 4 orders of magnitude
when AuNP-1 nanozyme was used. More, even compared with the unassembled
catalytic molecule Zn-1 (i.e., the complex of 1,4,7-triazacyclononane (TACN) and
zinc ion), AuNP-1 nanozyme exhibited more than 600 times rate acceleration. It
should note that the AuNP-1 nanozyme catalyzed reaction also obeyed a typical
Michaelis-Menten kinetics (Fig. 3.2d) [2, 5]. When alkanethiol terminated with
ammonium was used to modify the AuNPs, the formed AuNP-based complexes
were practically inactive. This control experiment confirmed that the catalytic
activity of AuNP-1 nanozyme was from the assembled monolayers rather than the
AuNP core. The AuNP-1 nanozyme could also catalyze the cleavage of RNA
dinucleotides (such as ApA, CpC, and UpU) [2].
Detailed studies revealed that the superior catalytic activity of AuNP-1 nano-
zyme could be attributed to its several unique features [2, 5]. First, it has been
showed that the effective local concentration of guest molecules (e.g., the catalytic
substrate HPNPP in this case) could be significantly enhanced due to both the
electrostatic and hydrophobic interactions between the positively charged

Fig. 3.2 a Comparison of catalyzed and uncatalyzed transphosphorylation of HPNPP. b Structure


of Zn-1 complex. c Proposed catalytic pocket of AuNP-1 nanozyme. d Michaelis-Menten
saturation kinetics of the transphosphorylation catalyzed by AuNP-1. E Dependence of the rate
constant for the transphosphorylation of HPNPP with AuNP-1 on Zn ion concentration. c and
d Adapted from Ref. [5], Copyright 2015, with permission from American Chemical Society.
e Adapted from Ref. [2], Copyright 2004, with permission from John Wiley and Sons
34 3 Metal-Based Nanomaterials for Nanozymes

monolayer and the guest molecules [5, 6]. Second, the assembled catalytic moieties
(i.e., the complex of TACN-Zn2+) were multivalent and exhibited cooperative
behavior toward the catalysis. The cooperativity was indicated by the sigmoidal
curve for zinc ion concentration-dependent catalytic activities (Fig. 3.2e) [2, 5]. The
cooperative effect was further validated by experiments and theoretical analysis [7].
It suggested that two neighboring catalytic moieties (i.e., two TACN-Zn2+) were
required to form a catalytic pocket for the cooperative catalysis (Fig. 3.2c) [2, 5, 7].
Such a catalytic pocket mimicked the ones in natural enzymes. Third, a 0.4 unit of
pKa decrease of the assembled TACN-Zn2+ complexes due to the close proximity
effect may also play a role in the enhanced activity [2]. Fourth, the strong Au-S
interaction made the self-assembly very easy to be carried out. Also it endowed the
assembled AuNP-based catalytic complexes with much higher stability compared
with micelle-based systems [5]. Moreover, its catalytic activity could be rationally
regulated. Since the catalytic activity of AuNP-1 nanozyme was from the com-
plexed Zn2+, the activity could be reversibly switched off by using a Zn2+ chelating
reagent and could be subsequently restored by re-adding Zn2+ [8].
For natural metallonucleases, their active sites have low dielectric constant (ε).
As shown in Eq. 3.1 below, the electrostatic interaction between the enzyme and its
substrate would be favored at active sites of low dielectric constant. To mimic such
a unique microenvironment, low polarity should be introduced into the active sites.

Eelectrostatic / ðQ1  Q2 Þ=ðe  r1;2 Þ ð3:1Þ

To test this hypothesis, AuNP-based nanozymes with different polarities have


been prepared and their metallonuclease mimicking activities have been studied [9].
As shown in Fig. 3.3, AuNP-3 and AuNP-5 with lower polarities exhibited higher
activities while AuNP-2 and AuNP-4 with higher polarities exhibited lower
activities. Therefore, the catalytic activities were well correlated with the polarities

Fig. 3.3 a AuNP-based nanozymes with different polarities. b Transphosphorylation of HPNPP


with AuNP-based nanozymes, highlighting the dianionic transition state. c Rate of HPNPP
cleavage with different nanozymes. Adapted from Ref. [9], Copyright 2014, with permission from
American Chemical Society
3.1 Metal Nanomaterials with Catalytic Monolayers (Type I) 35

of the assembled monolayers. The lowered polarities would enhance the electro-
static interaction between the dianionic transition state and the active sites of the
nanozymes (i.e., Zn2+ complex), which therefore improved the catalytic activities.
This study also suggested that catalytic activities of the nanozymes could be tuned
by modulating the monolayers assembled.
(b) Alkanethiol-protected AuNPs as RNase mimics: applications
The above-mentioned RNA mimics have been explored for numerous interesting
applications [10–12]. For instance, colorimetric detection of natural enzyme activity
through catalytic signal amplification with AuNP-1 nanozyme has been reported
[10]. As shown in Fig. 3.4, a natural enzyme substrate (i.e., a peptide in this study)
would interact strongly with AuNP-1 nanozyme and thus inhibit its RNase mim-
icking activity due to the shielding effect. The presence of the natural enzyme
would cleave the substrate into smaller fragments and thus significantly weaken the
interactions with AuNP-1 nanozyme. Therefore, the cleaved fragments would be
released from AuNP-1 nanozyme and thus restore the catalytic activity of AuNP-1
nanozyme. The restored AuNP-1 nanozyme would then convert HPNPP into the
colored product p-nitrophenol, which was used as a reporter. Through the cascade
amplification, the natural enzyme could be determined with higher sensitivity and
selectivity. Moreover, this design was a general approach, which has been extended
to other enzymes by using the corresponding peptide substrates [10].
A sensitive and selective strategy for Hg2+ detection has been demonstrated [11].
AuNP-1 nanozyme had stronger affinity toward a fluorescent reporter than thy-
midine probes (i.e., TDP, TMP, or cTMP), which resulted in the fluorescence
quenching of the reporter. In the presence of Hg2+, the probes would form T-Hg2+-
T complex and the complex could compete against the fluorescent reporter for
binding onto AuNP-1 nanozyme. Therefore, the reporter would be released from
AuNP-1 nanozyme and its fluorescence was recovered. Hg2+ at nanomolar

Fig. 3.4 Detection of enzyme activity through catalytic signal amplification with AuNP-1
nanozyme. Reprinted from Ref. [10], Copyright 2011, with permission from John Wiley and Sons
36 3 Metal-Based Nanomaterials for Nanozymes

Fig. 3.5 a Structure of BAPA. b Structure of AuNP-6 nanozyme. c Catalytic cleavage of BNP
with DNase mimic. Adapted from Ref. [13], Copyright 2008, with permission from American
Chemical Society

concentrations has been successfully detected with the proposed sensing strategy. It
has also been demonstrated that AuNP-1 nanozyme could be used to mimic the key
features of cellular signaling pathways [12].
(c) Alkanethiol-protected AuNPs as other enzyme mimics
It is relatively straightforward to design other enzyme mimics by changing the
terminal TACN moiety of the alkanethiol chain to other functional moieties [3, 4,
13–23].
For instance, a DNase mimic has been developed by assembling BAPA termi-
nated alkanethiol onto AuNPs (Fig. 3.5) [13]. The cleavage of BNP (bis-p-nitro-
phenyl phosphate), a DNA model substrate, into MNP (p-nitrophenyl phosphate)
and p-nitrophenolate has been accelerated with AuNP-6 nanozyme by 300,000-
folds over the background cleavage reaction. Moreover, DNA molecules
(such as pBR 322 plasmid DNA) have been successfully cleaved with AuNP-6
nanozyme [13].
DNase could also be mimicked by using AuNPs modified with Ce(IV) complex
terminated alkanethiol monolayers [15]. As many as 2.5 million-fold rate acceler-
ation for the BNP cleavage was observed using the AuNPs-Ce(IV)-based nano-
zymes. The superior catalytic activity was also attributed to the unique features of
the nanozyme, such as the cooperative catalysis.
(d) RNase (DNase) mimics using other supporting cores
Since AuNPs mainly acted as the supporting core, they could be replaced by
other supporting materials. Several studies have showed that by assembling the
catalytic moieties (such as TACN-Zn2+ complex) onto dendrimer, silica, etc.,
RNase and DNase mimics could be obtained (Fig. 3.6) [24–27]. Compared with the
facile Au-S chemistry, the synthesis and purification of these enzyme mimics are
much more difficult and time consuming. Other noble metal nanoparticles (such as
AgNPs), in principle, could also be employed as the supporting core. However, no
study has been reported, which might be due to the easy oxidation of AgNPs.
3.1 Metal Nanomaterials with Catalytic Monolayers (Type I) 37

Fig. 3.6 Dendrimer (a), silica particles (b), and resin (c) as inert supporting materials for catalytic
monolayers. a Adapted from Ref. [24], Copyright 2007, with permission from American Chemical
Society; b Adapted from Ref. [25], Copyright 2009, with permission from Royal Society of
Chemistry; c Adapted from Ref. [26], Copyright 2009, with permission from Elsevier

3.1.2 AuNPs Protected by Alkanethiol with Non-covalently


Assembled Catalytic Moieties

A few studies have showed that the catalytic moieties could be non-covalently
assembled onto (or into) the alkanethiol-protected AuNPs [28–30]. As shown in
Fig. 3.7a, the peptide ligation has been promoted when the two peptide fragments
were electrostatically assembled onto trimethylammonium-functionalized AuNPs
[28]. The transesterification of the p-nitrophenyl ester of N-carboxybenzyl-
phenylalanine was also significantly enhanced by two orders of magnitude when
both the substrate and the catalytic peptide were non-covalently assembled onto
trimethylammonium functionalized AuNPs (Fig. 3.7b) [29]. For both of the
examples, the non-covalent assembly was mainly driven by the electrostatic (and
hydrophobic) interactions. The prominent catalytic activities were owing to the
close proximity of the substrate and the catalyst as well as the unique microenvi-
ronment [28, 29].
Rotello and co-workers developed AuNPs-based nanozymes by encapsulating
hydrophobic molecular catalysts (e.g., Ru complex and Pd complex) within the
hydrophobic region of alkanethiol monolayers on AuNPs cores (Fig. 3.8a) [30]. To
protect the encapsulated molecular catalysts from release, cucurbit[7]uril (CB[7])
was used to cap the head groups of the alkanethiol monolayers. When capped by
CB[7], the nanozymes were inactive. The presence of a competing guest
(i.e., 1-adamantylamine (ADA)) could bind onto CB[7] and form supramolecular
complexes. Therefore, the molecular catalysts were exposed to the surrounding
environment and the nanozymes were subsequently activated. Based on such as
38 3 Metal-Based Nanomaterials for Nanozymes

Fig. 3.7 Peptide ligation (a) and transesterification (b) catalyzed by catalysts non-covalently
assembled onto AuNPs. a Reprinted from Ref. [28], Copyright 2007, with permission from
American Chemical Society; b Adapted from Ref. [29], Copyright 2012, with permission from
American Chemical Society

supramolecular regulation mechanism, the activities of the nanozymes were mod-


ulated in vivo in living cells. As shown in Fig. 3.8, the uptaken nanozymes could
activate nonfluorescent dye for cellular imaging. Moreover, the nanozyme could be
used to activate pro-drug (i.e., pro-5FU) in living cells. With such in situ activation

Fig. 3.8 a Schematic of the nanozyme, consisting of an AuNP core and an alkanethiol monolayer
with encapsulated catalysts. b Cellular uptake of the nanozyme and its regulation. c Activation of
nonfluorescent dye with the nanozyme. d Activation of pro-drug dye with the nanozyme. Adapted
from Ref. [30], Copyright 2015, with permission from Nature Publishing Group
3.1 Metal Nanomaterials with Catalytic Monolayers (Type I) 39

strategy, the toxic side effects of 5FU could be effectively minimized. This study
may provide a facile strategy for tuning the activities of nanozymes in living cells.

3.1.3 AuNPs Protected by Thiolated Biomolecules

Other AuNPs-based nanozymes have been developed by assembling thiolated


biomolecules on the AuNP cores [31–34]. Cao et al. has designed a nanozyme to
mimic the function of RNA-induced silencing complex (RISC) machinery [31].
With the guidance of a regulatory ssRNA, a natural RISC would bind to the
complementary target ssRNA and then silence its function by inducing target RNA
cleavage with a nuclease contained within the RISC. To mimic the features of a
RISC, both nucleases and regulatory ssRNA have been co-assembled onto an
AuNP core (Fig. 3.9a). In the presence of a target ssRNA, the assembled regulatory
ssRNA would form a duplex with the target ssRNA and thus bring it close to the
assembled nuclease for cleavage. To demonstrate the hypothesis, the anti-HCV
(hepatitis C virus) efficacy of the nanozyme was evaluated using a HCV replicon

Fig. 3.9 a Schematic of the nanozyme for RNA silencing. b Anti-HCV effects of the anti-HCV
nanozyme in FL-Neo cells. Adapted from Ref. [31], Copyright 2012, with permission from
National Academy of Sciences
40 3 Metal-Based Nanomaterials for Nanozymes

cell culture system (i.e., an FL-Neo cell line). As shown in Fig. 3.9b, the HCV
replication has been successfully inhibited with the nanozyme treatment. The
in vivo efficacy of the nanozyme was then evaluated using a xenotransplantation
mouse model. Remarkably, the nanozyme treatment resulted in more than 99 %
decrease of HCV RNA. Considering the non-detectable cellular interferon response,
the designed nanozyme may be used as a potent nanomedicine for viral infections
and cancers [31]. Other alternative approaches were developed for RNA
interference-independent gene regulation [32, 33, 35, 36].

3.2 Metal Nanomaterials with Intrinsic Enzyme


Mimicking Activities (Type II)

In 2004, the intrinsic oxidase mimicking activity of “naked” citrate-coated AuNPs


was reported [37], since then metal nanomaterials with intrinsic enzyme mimicking
activities have been extensively studied [38–64]. It has been showed that metal
nanomaterials (such as Au, Ag, Pd, Pt, etc.) could mimic oxidase, SOD, catalase,
peroxidase, etc. These nanozymes are unique in several aspects. First, most of them
have multiple enzyme mimicking activities, which are dependent on their
microenvironments. For instance, AuNPs exhibited catalase mimicking activity at
high pH while they showed SOD-like activity at low pH [47]. Second, their enzyme
mimicking activities could be tuned by forming alloys with other metals or by
exposing specific facets [65]. Third, the catalytic activities could also be enhanced
by exploring the plasmonic properties of noble metal nanomaterials [65].

3.2.1 Metal Nanomaterials as GOx Mimics

Glucose oxidase (GOx) catalyzes the oxidation of glucose into gluconic acid
(gluconate) and H2O2 with oxygen. Rossi and co-workers reported the selective
oxidation of glucose with oxygen using AuNPs load on carbon support as the
catalyst [66]. The high selectivity was attributed to the two facts: first, it has
successfully avoided the isomerization of glucose to fructose under acidic condi-
tions; second, it has enhanced the oxidation of the aldehydic group of glucose but
did not affect the oxidation of the alcoholic group [66].
Later they discovered that even citrate-coated AuNPs of 3–6 nm could catalyze
the aerobic oxidation of glucose with dissolved oxygen [37, 39, 40, 67]. The
AuNPs catalyzed reaction was similar to the GOx catalyzed one in terms of the
production of H2O2 and gluconate. Therefore, the AuNPs could be regarded as
GOx mimics. Detailed mechanism studies suggested that the catalytic reaction
followed an Eley–Rideal mechanism (Fig. 3.10) [39, 40, 67]. Briefly, a hydrated
glucose anion first interacted with Au atoms onto the AuNPs’ surface, which would
3.2 Metal Nanomaterials with Intrinsic Enzyme Mimicking Activities (Type II) 41

Fig. 3.10 Proposed molecular mechanism for the GOx-like activity of AuNPs. Reprinted from
Ref. [67], Copyright 2006, with permission from John Wiley and Sons

lead to electron-rich Au species. Then by reacting with the electron-rich Au species


and forming a dioxogold intermediate, oxygen was activated. Two electrons then
transferred from glucose to oxygen via the dioxogold intermediate, producing H2O2
and gluconate [67]. The AuNPs catalyzed reaction also obeyed Michaelis–Menten
kinetics [39, 40]. It also showed that the catalytic activities of the AuNPs were
inversely proportional to their size, further confirming their intrinsic enzyme
mimicking properties [37].
Since Rossi’s work, numerous groups have studied the GOx mimicking activ-
ities of various AuNPs [38, 48–51, 68–83]. For instance, it showed that the nature
of supports played a critical role in the catalytic activities of supported AuNPs as
GOx mimics [50]. Besides AuNPs-based GOx mimics, a few other metal nano-
materials have also been explored for mimicking GOx [52–54]. For example, Pd
nanoparticles on γ-Al2O3 support exhibited good selectivity toward glucose oxi-
dation [52]. It has also demonstrated that Au atom decorated Pd nanoparticles
exhibited higher catalytic activity toward glucose oxidation when compared with
AuNPs, PdNPs, and Pd/Au alloys [54]. It further demonstrated that Au-containing
bimetallic and trimetallic nanoparticles showed higher catalytic activities than
monometallic ones [53]. The enhanced activities have been ascribed to several
factors, such as electronic charge transfer effect among different metals, geometric
effect, and structural changes [53].

3.2.2 Metal Nanomaterials as Multiple Enzyme Mimics

As mentioned above, many metal nanomaterials have exhibited multiple enzyme


mimicking activities under different conditions [47, 65, 84–89]. For example,
PtNPs encapsulated within ferritin have exhibited both peroxidase and catalase
mimicking activities [84]. Due to the protection of the ferritin shell, the nanozymes
showed high stability. The peroxidase mimicking activity was confirmed by oxi-
dizing colorless substrates (i.e., TMB and DAB) with H2O2 into the corresponding
42 3 Metal-Based Nanomaterials for Nanozymes

colored products in the presence of the nanozymes. The catalase mimicking activity
was established by producing oxygen bubbles from H2O2 decomposition and by
inhibiting the formation of hydroxyl radicals from H2O2 in the presence of the
nanozymes. Moreover, they exhibited peroxidase-like activity at slightly acidic
condition and physiological temperature. When the pH and reaction temperature
increased, the catalase-like activity dominated [84]. The SOD and oxidase mim-
icking activities have also been observed in several nanomaterials [47, 86–88].
To better understand the multiple enzyme mimicking activities of metal nano-
materials, Wu, Gao and co-workers have carried out detailed computational studies
[65, 89]. Au, Ag, Pt, and Pd nanomaterials exhibited pH-switchable peroxidase-like
and catalase-like activities (i.e., they were peroxidase and catalase mimics at acidic
and basic conditions, respectively) [89]. To unravel the origin of such
pH-switchable catalytic activities, the H2O2 adsorption and the H2O2 decomposi-
tion behaviors on metal surfaces were calculated. First, the calculation of H2O2
adsorption on Au(111) surface indicated that H2O did not prevent H2O2 adsorption
at neutral pH. At acidic pH, H would be pre-adsorbed, which in turn weakened the
H2O2 adsorption on Au slightly. At basic pH, H2O2 interacted with Au atoms in the
vicinities of the pre-adsorbed OH. As shown in Fig. 3.11a, it favored the base-like
decomposition pathway at neutral conditions, which would produce adsorbed H2O
(i.e., adsorbed H2O*) and adsorbed O (i.e., adsorbed O*). However, the adsorbed O
may not form O2 due to the high energy barrier of 1.42 eV. As shown in
Fig. 3.11B, at acidic conditions, it still preferred the base-like decomposition
pathway. Due to the presence of pre-adsorbed H, H2O2 would be decomposed into
an adsorbed H2O (i.e., adsorbed H2O*) and an adsorbed OH (i.e., OH*). The OH*
would be converted into H2O* and O*. The formed O* subsequently oxidized
organic substrates by abstracting H atom from them. Therefore, the Au(111) surface
would act as a peroxidase mimic at acidic conditions. On the other hand, at basic
conditions, it still preferred the acid-like decomposition pathway (Fig. 3.11c). In
this case, one H from H2O2 would be react with the pre-adsorbed OH to form H2O*
and HO*2. The newly formed HO*2 would transfer one H to another H2O2, resulting
the formation of H2O* and O*2. Therefore, the Au(111) surface would act as a
catalase mimic at basic conditions. The above-calculated results were consistent
with the experimental results [89].
Further calculation revealed that for both Au(110) and Au(211) surfaces, they
also favored base-like decomposition and acid-like decomposition pathway at
acidic and basic conditions, respectively. In other words, both of them acted as
peroxidase and catalase mimics at acidic and basic conditions, respectively. Since
the enzyme mimicking activities were not dependent on the specific surfaces, it
suggested that the catalytic activities were intrinsic properties. However, among the
three surfaces, Au(111) surface exhibited the least catalytic activities due to the
highest energy barriers [89].
To understand the different enzyme mimicking activities of different metal
nanomaterials, the H2O2 decomposition on Ag(111), Pt(111), and Pd(111) surfaces
was also calculated [89]. Several conclusions could be obtained from the calcula-
tion results. First, all of them followed the same reactions pathways as Au(111)
3.2 Metal Nanomaterials with Intrinsic Enzyme Mimicking Activities (Type II) 43

Fig. 3.11 pH-switchable enzyme-mimic activities of metals. Calculated reaction energy profiles
for H2O2 decomposition on the Au(111) surface in neutral (a), acidic (b) and basic (c) conditions
(unit: eV). d Relationships between adsorption energies (Eads) and activation energies (Eact) for
H2O2 decompositions on metal’s (111) surfaces at acidic (left) and basic (right) conditions. e TEM
images of the nanorods. Enzyme-mimic activities of Au@Pd nanorods (0.1 nM) at 0.1 m PBS
buffers with different pH values for peroxidase with f 20 mm H2O2 + 0.4 mm OPD and for
catalase with h 20 mm H2O2, respectively. Peroxidase-like activity at pH = 4.5 (g) and
catalase-like activity at pH = 7.4 (i) for the metals. Unless indicated, reaction temperature in (f–i))
is 30 °C. Reprinted from Ref. [89], Copyright 2015, with permission from Elsevier

surface did due to their structure similarities (i.e., they acted as peroxidase and
catalase mimics at acidic and basic conditions, respectively). Second, the activation
energies (Eact) of them for both mimics followed the order: Au(111) < Ag
(111) < Pt(111) < Pd(111). Third, the adsorption energies (Eads) and activation
energies (Eact) exhibited an approximate linear relationship (Fig. 3.11d). It also
suggested that the affinities of the metal surface toward H2O2 followed the order:
Au(111) < Ag(111) < Pt(111) < Pd(111). Therefore, Eads could be practically used
to estimate the relative catalytic activities of nanozymes [89].
44 3 Metal-Based Nanomaterials for Nanozymes

The above calculations indicated that for both mimics, their catalytic activities
would follow the order: Au(111) < Ag(111) < Pt(111) < Pd(111). To test the
hypothesis, four nanorods (i.e., Au, Au@Ag, Au@Pt, and Au@Pd) were synthe-
sized and their activities were evaluated (Fig. 3.11e). First, the pH-switchable
activities were observed for Au@Pd nanorods (Fig. 3.11f and 3.11h). Second, for
different metal nanomaterials, both enzyme mimicking activities followed the order:
Au(111), Ag(111) < Pt(111), Pd(111) (Fig. 3.11g and 3.11i), which was consistent
with the calculations. Note that, the following orders were observed: Ag(111) < Au
(111) and Pd(111) < Pt(111). Such deviation between the experimental and cal-
culation results could be attributed to the following facts. For Ag(111), it could be
easily oxidized and therefore affected its activities. For Pt(111), it had larger surface
compared with Pd(111). Considering these facts, the calculations have provided
mechanistic insights of the metals’ enzyme mimicking activities and have provided
a way to predict the nanozymes’ activities [89].
Wu, Gao and co-workers also studied the origins of oxidase and SOD mimicking
activities of Au, Ag, Pt, Pd, and their alloys [65]. For the oxidase mimicking
activities, the adsorption of O2 and breakage of O–O bond (i.e., the dissociation of
O2) should be critical since it would transfer the magnetic moments from 3O2 to the
metal and thus allow originally spin forbidden 3O2 react with its substrate via O*.
For Au(111), Ag(111), Pt(111), and Pd(111), though O2 could be adsorbed onto all
of them and subsequently weakened the O–O bond of O2 in an energy favorable
manner, only Pt(111) and Pd(111) exhibited low Eact and negative Er (reaction
energy) values. The high Eact and positive Er values for Au(111) and Ag(111)
suggested that Pt(111) and Pd(111) would exhibit oxidase mimicking activities
while Au(111) and Ag(111) would not. Since the oxidase mimicking activities of
AuNPs have been reported [37], there should be some factors responsible for such
activities. To further reveal the origin of the AuNPs’ oxidase-like activities, the
calculations for Au(110) and Au(211) surfaces were also carried out. Though Au
(110) still did not show favorable oxidase mimicking activity, Au(211) did owing to
the low Eact and negative Er. It suggested that the oxidase mimicking activities of
the reported AuNPs might be originated from their high-energy facets, such as Au
(211) [65]. Low-spin state 1O2 could be converted from 3O2 by the surface plasmon
resonance of noble metals. Though the dissociation of 1O2 were kinetically easier
than that of 3O2, the Eact values were still too large for Au(111) and Ag(111)
surfaces. Therefore, the conversion of O2 from 3O2 to 1O2 would not substantially
affect the metal-mediated O2 dissociation and thus their oxidase-like activities [65].
By forming alloys, the oxidase-like activities could be significantly modulated
[65]. The calculation results indicated that the formation of AuAg alloys has sig-
nificantly reduced the Eact values compared with Au and Ag themselves. Moreover,
Er values for the alloys also became negatively. Therefore, the formation of AuAg
alloys would enhance the oxidase mimicking activities of both Au and Ag. In
contrast, the formation of alloys between Pd (or Pt) and Au led to the increase of
Eact values and the positive Er values [65]. Therefore, the oxidase mimicking
activities were disfavored for the AuPd or AuPt alloys. These results suggested a
3.2 Metal Nanomaterials with Intrinsic Enzyme Mimicking Activities (Type II) 45

possible approach to modulating the metals’ oxidase mimicking activities (in both
positive and negative manners).
Since O∙− − ∙
2 radical can easily react with H2O to produce OH and HO2, the SOD
mimicking activities of metal nanomaterials are mainly due to their capability for
rearrange HO∙2 into H2O2 and O2. The calculation results showed that the Eact
values for the rearrangement on both Au(111) and Pt(111) were very small, indi-
cating that both of the metals could act as high efficient SOD mimics [65].

3.2.3 Applications

Along with exploring the mechanisms of the metal-based nanozymes, they have
also been widely used for various applications [79, 80, 86, 90–96].
(a) Immunoassays
The metal nanomaterials-based oxidase, peroxidase, and even catalase mimics
have been employed for immunoassays [86, 91, 95, 97–101]. For instance, sandwich
assays for interleukin 2 (IL-2, a cytokine) have been developed with Au@Pt
nanorods [86]. As shown in Fig. 3.12, both the oxidase and peroxidase mimicking
activities of the Au@Pt nanorods could be employed for detection. In both cases,
colorless TMB would be oxidized into colored products when IL-2 was present [86].
When the nanozymes were conjugated with tumor cell targeting molecules, the
conjugates could be used for tumor cell immunoassay. For instance, when Au
nanoclusters decorated graphene oxide was labeled with folic acid, the conjugates
showed peroxidase-like activity and have been used for selectively quantifying
MCF-7 tumor cells [98]. Using Pt nanoparticles as catalase mimics, a volumetric
bar-chart chip has been developed for detection of biomarkers both in serum and on
cell surface [99]. An integrin specific peptide was used to modify AuNPs-based
peroxidase mimics. The obtained conjugates were then used as specific probes for
cancer detection [101].
(b) Glucose and other bioactive small molecules detection
By combing GOx with nanomaterials-based peroxidase mimics, glucose could
be detected [58, 64, 87, 102–104]. Due to high specificity of natural GOx, these
assays exhibited high selectivity toward glucose detection. Since AuNPs could act
as GOx mimics, they were combined with hemin (a peroxidase mimic) together for
colorimetric detection of glucose [81]. The selectivity of AuNPs-based GOx mimics
remained to be studied.
In principle, when a natural oxidase was combined with nanomaterials-based
peroxidase mimics, the corresponding oxidase substrate as the target of interest
could be determined (also see Tables A1 and A2). For example, a colorimetric
assay for cholesterol was developed by combining cholesterol oxidase with Pt
nanoparticles-based peroxidase mimics [64]. More interestingly, sarcosine, a
46 3 Metal-Based Nanomaterials for Nanozymes

Fig. 3.12 Immunoassays for interleukin 2. a Au@Pt nanorods-based assays. b Detection of IL-2
using Au@Pt nanorods’ oxidase (i.e., without H2O2) or peroxidase (i.e., with H2O2) mimicking
activities. Reprinted from Ref. [86], Copyright 2011, with permission from Elsevier

potential biomarker for prostate cancer, in clinical samples has been successfully
determined by using sarcosine oxidase and Pd nanoparticles-based peroxidase
mimics [105].
(c) Detection of analytes by modulating nanozymes’ activities
The nanozymes’ activities could be modulated by changing their sizes, surface
coating, etc. Based on these phenomena, numerous strategies have been developed
for sensing analytes of interest [78, 106–113].
3.2 Metal Nanomaterials with Intrinsic Enzyme Mimicking Activities (Type II) 47

Fan, Li, and co-workers found that the catalytic activities of AuNPs-based GOx
mimics were controlled in a self-limited manner [107]. As discussed in Chap. 2,
ssDNA and dsDNA have different affinities toward AuNPs. By combing the
AuNPs’ GOx mimicking activities with DNA modulation, a general sensing plat-
form was developed (Fig. 3.13) [106]. The probe ssDNA would interact with
AuNPs nanozymes and thus inhibited their catalytic activities. The presence of
targets (such as a complementary ssDNA, a complementary miRNA, or an analyte
for an aptamer if the probe ssDNA was an aptamer), the probe ssDNA would form a
complex with its target and thus recover the activity of AuNPs. The nanozymes
activities were then monitored with suitable output signals. For example, when it
was coupled with HRP, either colorimetric or chemiluminescent signal would be
produced. The intrinsic plasmonoic signals of the AuNPs could also be used for
signaling [106]. Inspired by this work, several reports showed that other functional
nucleic acids (including aptamers) could be used to modulate the nanozymes’
activities [108, 109, 111, 114, 115]. For example, a colorimetric method for the
detection of kanamycin has been developed [108]. Kanamycin is an aminogly-
coside antibiotic widely used in veterinary medicine. Its aptamer would bind onto
AuNPs and inhibit their peroxidase-like activities. The specific interaction between
kanamycin and its aptamer would recover the inhibited activities of the nanozymes.
Via such a switchable strategy, kanamycin has been detected with good sensitivity
and selectivity. Using the same principle, Li et al. reported the colorimetric
detection of ricin with its aptamer and peroxidase-like AuNPs [115].
Besides DNA, other biomolecules could also be used to modulate the nano-
zymes activities and to develop biosensors [78, 113, 116]. Li and co-workers found
that melamine could significantly improve the peroxidase-like activity of AuNPs
[116]. By making use of such enhancement effect, they proposed a colorimetric
method for melamine detection (Fig. 3.14). The developed method was robust

Fig. 3.13 Illustration of the GOx-like catalytic activity of AuNPs regulated by DNA hybridiza-
tion, which can be either amplified by HRP cascaded color or chemiluminescence variations (path
a) or lead to nanoplasmonic changes owing to size enlargement (path b). Orange strand = target,
green strand = adsorption probe. Reprinted from Ref. [106], Copyright 2011, with permission
from John Wiley and Sons
48 3 Metal-Based Nanomaterials for Nanozymes

Fig. 3.14 Detection of melamine based on its enhancement toward the peroxidase-like activity of
AuNPs. Reprinted from Ref. [116], Copyright 2014, with permission from Elsevier

Fig. 3.15 Detection of Hg2+ based on its inhibition toward the peroxidase-like activity of PtNPs.
Reprinted from Ref. [117], Copyright 2015, with permission from Elsevier

enough that it has been employed to detect spiked melamine in raw milk and milk
powder [116].
Metal ions have also been used to modulate the nanozymes’ activities [117,
118]. As shown in Fig. 3.15, the peroxidase-like activity of PtNPs could be
selectively inhibited by Hg2+ due to the specific aurophilic/metallophilic interac-
tions between Hg2+ and Pt0 [117]. Based on such specific inhibition, Hg2+ has been
selectively detected against several other metal ions (such as Na+, Mg2+, Ca2+,
Mn2+, Ni2+, Zn2+, Co2+, Cu2+, Pd2+, Cd2+, Fe3+, and Au3+).
(d) Other applications
Other potential applications of metal nanomaterials-based nanozymes have also
been explored [61, 119]. For instance, Qu, Ren, and co-workers have studied the
antibacterial activity of AuNPs encapsulated within mesoporous silica [119]. The
nanozymes exhibited both oxidase and peroxidase mimicking activities. Since ROS
(reactive oxygen species) species (such as O∙− ∙
2 and HO ) were involved in the
mimicking activities, it was reasonable to test the nanozymes’ antibacterial activity.
As expected, the nanozymes showed antibacterial properties against both
Gram-negative and Gram-positive bacteria [119].
References 49

References

1. Love, J. C., Estroff, L. A., Kriebel, J. K., Nuzzo, R. G., & Whitesides, G. M. (2005).
Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chemical
Reviews, 105, 1103–1169.
2. Manea, F., Houillon, F. B., Pasquato, L., & Scrimin, P. (2004). Nanozymes:
Gold-nanoparticle-based transphosphorylation catalysts. Angewandte Chemie-International
Edition, 43, 6165–6169.
3. Levy, R. (2006). Peptide-capped gold nanoparticles: Towards artificial proteins.
ChemBioChem, 7, 1141–1145.
4. Mancin, F., Prins, L. J., & Scrimin, P. (2013). Catalysis on gold-nanoparticle-passivating
monolayers. Current Opinion in Colloid & Interface Science, 18, 61–69.
5. Prins, L. J. (2015). Emergence of complex chemistry on an organic monolayer. Accounts of
Chemical Research, 48, 1920–1928.
6. Pieters, G., Pezzato, C., & Prins, L. J. (2013). Controlling supramolecular complex formation
on the surface of a monolayer-protected gold nanoparticle in water. Langmuir, 29,
7180–7185.
7. Zaupa, G., Mora, C., Bonomi, R., Prins, L. J., & Scrimin, P. (2011). Catalytic self-assembled
monolayers on Au nanoparticles: The source of catalysis of a transphosphorylation reaction.
Chemistry-A European Journal, 17, 4879–4889.
8. Pieters, G., Pezzato, C., & Prins, L. J. (2012). Reversible control over the valency of a
nanoparticle-based supramolecular system. Journal of the American Chemical Society, 134,
15289–15292.
9. Diez-Castellnou, M., Mancin, F., & Scrimin, P. (2014). Efficient phosphodiester cleaving
nanozymes resulting from multivalency and local medium polarity control. Journal of the
American Chemical Society, 136, 1158–1161.
10. Bonomi, R., Cazzolaro, A., Sansone, A., Scrimin, P., & Prins, L. J. (2011). Detection of
enzyme activity through catalytic signal amplification with functionalized gold nanoparticles.
Angewandte Chemie-International Edition, 50, 2307–2312.
11. Maiti, S., Pezzato, C., Martin, S. G., & Prins, L. J. (2014). Multivalent interactions regulate
signal transduction in a self-assembled Hg2+ sensor. Journal of the American Chemical
Society, 136, 11288–11291.
12. Pezzato, C., & Prins, L. J. (2015). Transient signal generation in a self-assembled
nanosystem fueled by ATP. Nature Communications, 6, 7790.
13. Bonomi, R., Selvestrel, F., Lombardo, V., Sissi, C., Polizzi, S., Mancin, F., et al. (2008).
Phosphate diester and DNA hydrolysis by a multivalent, nanoparticle-based catalyst. Journal
of the American Chemical Society, 130, 15744–15745.
14. Pengo, P., Baltzer, L., Pasquato, L., & Scrimin, P. (2007). Substrate modulation of the
activity of an artificial nanoesterase made of peptide-functionalized gold nanoparticles.
Angewandte Chemie-International Edition, 46, 400–404.
15. Bonomi, R., Scrimin, P., & Mancin, F. (2010). Phosphate diesters cleavage mediated by Ce
(IV) complexes self-assembled on gold nanoparticles. Organic & Biomolecular Chemistry, 8,
2622–2626.
16. Sheet, D., Halder, P., & Paine, T. K. (2013). Enhanced reactivity of a biomimetic iron(II)
alpha-keto acid complex through immobilization on functionalized gold nanoparticles.
Angewandte Chemie-International Edition, 52, 13314–13318.
17. Salvio, R., & Cincotti, A. (2014). Guanidine based self-assembled monolayers on Au
nanoparticles as artificial phosphodiesterases. RSC Advances, 4, 28678–28682.
18. Pieters, G., & Prins, L. J. (2012). Catalytic self-assembled monolayers on gold nanoparticles.
New Journal of Chemistry, 36, 1931–1939.
19. Kisailus, D., Najarian, M., Weaver, J. C., & Morse, D. E. (2005). Functionalized gold
nanoparticles mimic catalytic activity of a polysiloxane-synthesizing enzyme. Advanced
Materials, 17, 1234–1239.
50 3 Metal-Based Nanomaterials for Nanozymes

20. Yapar, S., Oikonomou, M., Veldersb, A. H., & Kubik, S. (2015). Dipeptide recognition in
water mediated by mixed monolayer protected gold nanoparticles. Chemical
Communications, 51, 14247–14250.
21. Li, X., Qi, Z., Liang, K., Bai, X., Xu, J., Liu, J., et al. (2008). An artificial supramolecular
nanozyme based on beta-cyclodextrin-modified gold nanoparticles. Catalysis Letters, 124,
413–417.
22. Zhang, Z., Fu, Q., Huang, X., Xu, J., Liu, J., & Shen, J. (2009). Construction of the active
site of metalloenzyme on au nc micelles. Chinese Journal of Chemistry, 27, 1215–1220.
23. Zhang, Z., Fu, Q., Li, X., Huang, X., Xu, J., Shen, J., et al. (2009). Self-assembled gold
nanocrystal micelles act as an excellent artificial nanozyme with ribonuclease activity.
Journal of Biological Inorganic Chemistry, 14, 653–662.
24. Martin, M., Manea, F., Fiammengo, R., Prins, L. J., Pasquato, L., & Scrimin, P. (2007).
Metallodendrimers as transphosphorylation catalysts. Journal of the American Chemical
Society, 129, 6982–6983.
25. Bodsgard, B. R., Clark, R. W., Ehrbar, A. W., Burstyn, J. N. (2009). Silica-bound copper(II)
triazacyclononane as a phosphate esterase: Effect of linker length and surface
hydrophobicity. Dalton Transactions, 2365–2373.
26. Zaupa, G., Prins, L. J., & Scrimin, P. (2009). Resin-supported catalytic dendrimers as
multivalent artificial metallonucleases. Bioorganic & Medicinal Chemistry Letters, 19,
3816–3820.
27. Savelli, C., & Salvio, R. (2015). Guanidine-based polymer brushes grafted onto silica
nanoparticles as efficient artificial phosphodiesterases. Chemistry-A European Journal, 21,
5856–5863.
28. Fillon, Y., Verma, A., Ghosh, P., Ernenwein, D., Rotello, V. M., & Chmielewski, J. (2007).
Peptide ligation catalyzed by functionalized gold nanoparticles. Journal of the American
Chemical Society, 129, 6676–6677.
29. Zaramella, D., Scrimin, P., & Prins, L. J. (2012). Self-assembly of a catalytic multivalent
peptide-nanoparticle complex. Journal of the American Chemical Society, 134, 8396–8399.
30. Tonga, G. Y., Jeong, Y., Duncan, B., Mizuhara, T., Mout, R., Das, R., et al. (2015).
Supramolecular regulation of bioorthogonal catalysis in cells using nanoparticle-embedded
transition metal catalysts. Nature Chemistry, 7, 597–603.
31. Wang, Z. L., Liu, H. Y., Yang, S. H., Wang, T., Liu, C., & Cao, Y. C. (2012).
Nanoparticle-based artificial RNA silencing machinery for antiviral therapy. Proceedings of
the National Academy of Sciences of the United States of America, 109, 12387–12392.
32. Yehl, K., Joshi, J. R., Greene, B. L., Dyer, R. B., Nahta, R., & Salaita, K. (2012). Catalytic
deoxyribozyme-modified nanoparticles for RNAi-independent gene regulation. ACS Nano, 6,
9150–9157.
33. Patel, S., Jung, D., Yin, P. T., Carlton, P., Yamamoto, M., Bando, T., et al. (2014).
Nanoscript: A nanoparticle-based artificial transcription factor for effective gene regulation.
ACS Nano, 8, 8959–8967.
34. Patel, S., Yin, P. T., Sugiyama, H., & Lee, K.-B. (2015). Inducing stem cell myogenesis
using nanoscript. ACS Nano, 9, 6909–6917.
35. Wang, Z., Wang, Z., Liu, D., Yan, X., Wang, F., Niu, G., et al. (2014). Biomimetic
RNA-silencing nanocomplexes: Overcoming multidrug resistance in cancer cells.
Angewandte Chemie-International Edition, 53, 1997–2001.
36. Rouge, J. L., Sita, T. L., Hao, L., Kouri, F. M., Briley, W. E., Stegh, A. H., et al. (2015).
Ribozyme-spherical nucleic acids. Journal of the American Chemical Society, 137,
10528–10531.
37. Comotti, M., Della Pina, C., Matarrese, R., Rossi, M. (2004). The catalytic activity of
“naked” gold particles. Angewandte Chemie-International Edition, 43, 5812–5815.
38. Wei, H., & Wang, E. K. (2013). Nanomaterials with enzyme-like characteristics
(nanozymes): Next-generation artificial enzymes. Chemical Society Reviews, 42, 6060–6093.
39. Beltrame, P., Comotti, M., Della Pina, C., & Rossi, M. (2004). Aerobic oxidation of glucose
I. Enzymatic catalysis. Journal of Catalysis, 228, 282–287.
References 51

40. Beltrame, P., Comotti, M., Della Pina, C., Rossi, M. (2006) Aerobic oxidation of glucose II.
catalysis by colloidal gold. Applied Catalysis A-General, 297, 1–7.
41. Lin, Y. H., Ren, J. S., & Qu, X. G. (2014). Nano-gold as artificial enzymes: Hidden talents.
Advanced Materials, 26, 4200–4217.
42. Della Pina, C., Falletta, E., Rossi, M. (2012). Update on selective oxidation using gold.
Chemical Society Reviews, 41, 350–369.
43. Tan, X., Deng, W., Liu, M., Zhang, Q., Wang, Y. (2009). Carbon nanotube-supported gold
nanoparticles as efficient catalysts for selective oxidation of cellobiose into gluconic acid in
aqueous medium. Chemical Communications, 7179–7181.
44. Zhang, J., Liu, X., Hedhili, M. N., Zhu, Y., & Han, Y. (2011). Highly selective and complete
conversion of cellobiose to gluconic acid over Au/Cs2HPW12O40 nanocomposite catalyst.
ChemCatChem, 3, 1294–1298.
45. An, D., Ye, A., Deng, W., Zhang, Q., & Wang, Y. (2012). Selective conversion of cellobiose
and cellulose into gluconic acid in water in the presence of oxygen, catalyzed by
polyoxometalate-supported gold nanoparticles. Chemistry-A European Journal, 18,
2938–2947.
46. Lee, L.-C., & Zhao, Y. (2014). Metalloenzyme-mimicking supramolecular catalyst for highly
active and selective intramolecular alkyne carboxylation. Journal of the American Chemical
Society, 136, 5579–5582.
47. He, W. W., Zhou, Y. T., Warner, W. G., Hu, X. N., Wu, X. C., Zheng, Z., et al. (2013).
Intrinsic catalytic activity of Au nanoparticles with respect to hydrogen peroxide
decomposition and superoxide scavenging. Biomaterials, 34, 765–773.
48. Miedziak, P. J., Alshammari, H., Kondrat, S. A., Clarke, T. J., Davies, T. E., Morad, M.,
et al. (2014). Base-free glucose oxidation using air with supported gold catalysts. Green
Chemistry, 16, 3132–3141.
49. Wang, Y., Van de Vyver, S., Sharma, K. K., & Roman-Leshkov, Y. (2014). Insights into the
stability of gold nanoparticles supported on metal oxides for the base-free oxidation of
glucose to gluconic acid. Green Chemistry, 16, 719–726.
50. Ishida, T., Kinoshita, N., Okatsu, H., Akita, T., Takei, T., & Haruta, M. (2008). Influence of
the support and the size of gold clusters on catalytic activity for glucose oxidation.
Angewandte Chemie-International Edition, 47, 9265–9268.
51. Zhan, P., Wang, Z.-G., Li, N., & Ding, B. (2015). Engineering gold nanoparticles with DNA
ligands for selective catalytic oxidation of chiral substrates. ACS Catalysis, 5, 1489–1498.
52. Liang, X., Liu, C.-J., & Kuai, P. (2008). Selective oxidation of glucose to gluconic acid over
argon plasma reduced Pd/Al2O3. Green Chemistry, 10, 1318–1322.
53. Zhang, H., & Toshima, N. (2013). Glucose oxidation using Au-containing bimetallic and
trimetallic nanoparticles. Catalysis Science & Technology, 3, 268–278.
54. Zhang, H., Watanabe, T., Okumura, M., Haruta, M., & Toshima, N. (2012). Catalytically
highly active top gold atom on palladium nanocluster. Nature Materials, 11, 49–52.
55. Lin, Y. H., Zhao, A. D., Tao, Y., Ren, J. S., & Qu, X. G. (2013). Ionic liquid as an efficient
modulator on artificial enzyme system: Toward the realization of high-temperature catalytic
reactions. Journal of the American Chemical Society, 135, 4207–4210.
56. Liu, Y., Purich, D. L., Wu, C. C., Wu, Y., Chen, T., Cui, C., et al. (2015). Ionic
functionalization of hydrophobic colloidal nanoparticles to form ionic nanoparticles with
enzyme like properties. Journal of the American Chemical Society, 137, 14952–14958.
57. Wang, S., Chen, W., Liu, A. L., Hong, L., Deng, H. H., & Lin, X. H. (2012). Comparison of
the peroxidase-like activity of unmodified, amino-modified, and citrate-capped gold
nanoparticles. ChemPhysChem, 13, 1199–1204.
58. Jv, Y., Li, B. X., & Cao, R. (2010). Positively-charged gold nanoparticles as peroxidiase
mimic and their application in hydrogen peroxide and glucose detection. Chemical
Communications, 46, 8017–8019.
59. Jiang, H., Chen, Z. H., Cao, H. Y., & Huang, Y. M. (2012). Peroxidase-like activity of
chitosan stabilized silver nanoparticles for visual and colorimetric detection of glucose.
Analyst, 137, 5560–5564.
52 3 Metal-Based Nanomaterials for Nanozymes

60. Zhang, K., Hu, X. N., Liu, J. B., Yin, J. J., Hou, S. A., Wen, T., et al. (2011). Formation of
PdPt alloy nanodots on gold nanorods: Tuning oxidase-like activities via composition.
Langmuir, 27, 2796–2803.
61. Shibuya, S., Ozawa, Y., Watanabe, K., Izuo, N., Toda, T., Yokote, K., et al. (2014).
Palladium and platinum nanoparticles attenuate aging-like skin atrophy via antioxidant
activity in mice. PLoS ONE, 9, 109288.
62. Chen, Z., Zhao, C., Ju, E., Ji, H., Ren, J., Binks, B. P., et al. (2016). Design of surface-active
artificial enzyme particles to stabilize pickering emulsions for high-performance biphasic
biocatalysis. Advanced Materials, 28, 1682–1688.
63. Fu, Y., Zhao, X. Y., Zhang, J. L., & Li, W. (2014). DNA-based platinum nanozymes for
peroxidase mimetics. Journal of Physical Chemistry C, 118, 18116–18125.
64. Zhang, W., Liu, X. Y., Walsh, D., Yao, S. Y., Kou, Y., & Ma, D. (2012).
Caged-protein-confined bimetallic structural assemblies with mimetic peroxidase activity.
Small (Weinheim an der Bergstrasse, Germany), 8, 2948–2953.
65. Shen, X., Liu, W., Gao, X., Lu, Z., Wu, X., & Gao, X. (2015). Mechanisms of oxidase and
superoxide dismutation-like activities of gold, silver, platinum, and palladium, and their
alloys: A general way to the activation of molecular oxygen. Journal of the American
Chemical Society, 137, 15882–15891.
66. Biella, S., Prati, L., & Rossi, M. (2002). Selective oxidation of D-glucose on gold catalyst.
Journal of Catalysis, 206, 242–247.
67. Comotti, M., Della Pina, C., Falletta, E., & Rossi, M. (2006). Aerobic oxidation of glucose
with gold catalyst: Hydrogen peroxide as intermediate and reagent. Advanced Synthesis &
Catalysis, 348, 313–316.
68. Lang, N. J., Liu, B. W., & Liu, J. W. (2014). Characterization of glucose oxidation by gold
nanoparticles using nanoceria. Journal of Colloid and Interface Science, 428, 78–83.
69. Delidovich, I. V., Moroz, B. L., Taran, O. P., Gromov, N. V., Pyrjaev, P. A., Prosvirin, I. P.,
et al. (2013). Aerobic selective oxidation of glucose to gluconate catalyzed by Au/Al2O3 and
Au/C: Impact of the mass-transfer processes on the overall kinetics. Chemical Engineering
Journal, 223, 921–931.
70. Odrozek, K., Maresz, K., Koreniuk, A., Prusik, K., & Mrowiec-Bialon, J. (2014).
Amine-stabilized small gold nanoparticles supported on A1SBA-15 as effective catalysts
for aerobic glucose oxidation. Applied Catalysis A-General, 475, 203–210.
71. Zhang, H., Lu, L., Cao, Y., Du, S., Cheng, Z., & Zhang, S. (2014). Fabrication of
catalytically active Au/Pt/Pd trimetallic nanoparticles by rapid injection of NaBH4. Materials
Research Bulletin, 49, 393–398.
72. Ishida, T., Okamoto, S., Makiyama, R., & Haruta, M. (2009). Aerobic oxidation of glucose
and 1-phenylethanol over gold nanoparticles directly deposited on ion-exchange resins.
Applied Catalysis A-General, 353, 243–248.
73. Okatsu, H., Kinoshita, N., Akita, T., Ishida, T., & Haruta, M. (2009). Deposition of gold
nanoparticles on carbons for aerobic glucose oxidation. Applied Catalysis A-General, 369,
8–14.
74. Benko, T., Beck, A., Geszti, O., Katona, R., Tungler, A., Frey, K., et al. (2010). Selective
oxidation of glucose versus CO oxidation over supported gold catalysts. Applied Catalysis
A-General, 388, 31–36.
75. Shi, L., Liu, K., Zou, X., Yin, M., & Suo, Z. (2010). Selective oxidation of glucose in the
presence of PVP-protected colloidal gold solutions. Chinese Journal of Catalysis, 31,
661–665.
76. Ma, C., Xue, W., Li, J., Xing, W., & Hao, Z. (2013). Mesoporous carbon-confined Au
catalysts with superior activity for selective oxidation of glucose to gluconic acid. Green
Chemistry, 15, 1035–1041.
77. Yin, H., Zhou, C., Xu, C., Liu, P., Xu, X., & Ding, Y. (2008). Aerobic oxidation of
D-glucose on support-free nanoporous gold. Journal of Physical Chemistry C, 112,
9673–9678.
References 53

78. Wang, G. L., Jin, L. Y., Dong, Y. M., Wu, X. M., & Li, Z. J. (2015). Intrinsic enzyme
mimicking activity of gold nanoclusters upon visible light triggering and its application for
colorimetric trypsin detection. Biosensors & Bioelectronics, 64, 523–529.
79. Wang, L. H., Zeng, Y., Shen, A. G., Zhou, X. D., & Hu, J. M. (2015). Three dimensional
nano-assemblies of noble metal nanoparticle-infinite coordination polymers as specific
oxidase mimetics for degradation of methylene blue without adding any cosubstrate.
Chemical Communications, 51, 2052–2055.
80. Zhou, H., Han, T., Wei, Q., & Zhang, S. (2016). Efficient enhancement of
electrochemiluminescence from cadmium sulfide quantum dots by glucose oxidase
mimicking gold nanoparticles for highly sensitive assay of methyltransferase activity.
Analytical Chemistry, 88, 2976–2983.
81. Lin, Y. H., Wu, L., Huang, Y. Y., Ren, J. S., & Qu, X. G. (2015). Positional assembly of
hemin and gold nanoparticles in graphene-mesoporous silica nanohybrids for tandem
catalysis. Chemical Science, 6, 1272–1276.
82. Huang, Y., Lin, Y., Ran, X., Ren, J., Qu, X. (2016) Self-assembly and compartmentalization
of nanozymes in mesoporous silica-based nanoreactors. Chemistry (Weinheim an der
Bergstrasse, Germany), 22, 5705–5711.
83. Lin, Y. H., Li, Z. H., Chen, Z. W., Ren, J. S., & Qu, X. G. (2013). Mesoporous
silica-encapsulated gold nanoparticles as artificial enzymes for self-activated cascade
catalysis. Biomaterials, 34, 2600–2610.
84. Fan, J., Yin, J.-J., Ning, B., Wu, X., Hu, Y., Ferrari, M., et al. (2011). Direct evidence for
catalase and peroxidase activities of ferritin-platinum nanoparticles. Biomaterials, 32,
1611–1618.
85. He, W., Zhou, Y.-T., Wamer, W. G., Boudreau, M. D., & Yin, J.-J. (2012). Mechanisms of
the pH dependent generation of hydroxyl radicals and oxygen induced by Ag nanoparticles.
Biomaterials, 33, 7547–7555.
86. He, W. W., Liu, Y., Yuan, J. S., Yin, J. J., Wu, X. C., Hu, X. N., et al. (2011). Au@Pt
nanostructures as oxidase and peroxidase mimetics for use in immunoassays. Biomaterials,
32, 1139–1147.
87. Liu, J. B., Hu, X. N., Hou, S., Wen, T., Liu, W. Q., Zhu, X., et al. (2012). Au@Pt core/shell
nanorods with peroxidase- and ascorbate oxidase-like activities for improved detection of
glucose. Sensors and Actuators B-Chemical, 166, 708–714.
88. Hu, X. N., Saran, A., Hou, S., Wen, T., Ji, Y. L., Liu, W. Q., et al. (2013). Au@PtAg
core/shell nanorods: tailoring enzyme-like activities via alloying. RSC Advances, 3,
6095–6105.
89. Li, J. N., Liu, W. Q., Wu, X. C., & Gao, X. F. (2015). Mechanism of pH-switchable
peroxidase and catalase-like activities of gold, silver, platinum and palladium. Biomaterials,
48, 37–44.
90. Hu, D., Sheng, Z., Fang, S., Wang, Y., Gao, D., Zhang, P., et al. (2014). Folate
receptor-targeting gold nanoclusters as fluorescence enzyme mimetic nanoprobes for tumor
molecular colocalization diagnosis. Theranostics, 4, 142–153.
91. Wang, Z. F., Zheng, S., Cai, J., Wang, P., Feng, J., Yang, X., et al. (2013). Fluorescent
artificial enzyme-linked immunoassay system based on Pd/C nanocatalyst and fluorescent
chemodosimeter. Analytical Chemistry, 85, 11602–11609.
92. He, S. B., Wu, G. W., Deng, H. H., Liu, A. L., Lin, X. H., Xia, X. H., et al. (2014). Choline
and acetylcholine detection based on peroxidase-like activity and protein antifouling property
of platinum nanoparticles in bovine serum albumin scaffold. Biosensors & Bioelectronics,
62, 331–336.
93. Zheng, C., Zheng, A. X., Liu, B., Zhang, X. L., He, Y., Li, J., et al. (2014). One-pot
synthesized DNA-templated Ag/Pt bimetallic nanoclusters as peroxidase mimics for
colorimetric detection of thrombin. Chemical Communications, 50, 13103–13106.
94. Lin, X. Q., Deng, H. H., Wu, G. W., Peng, H. P., Liu, A. L., Lin, X. H., et al. (2015).
Platinum nanoparticles/graphene-oxide hybrid with excellent peroxidase-like activity and its
application for cysteine detection. Analyst, 140, 5251–5256.
54 3 Metal-Based Nanomaterials for Nanozymes

95. Xia, X. H., Zhang, J. T., Lu, N., Kim, M. J., Ghale, K., Xu, Y., et al. (2015). Pd-Ir core-shell
nanocubes: A type of highly efficient and versatile peroxidase mimic. ACS Nano, 9,
9994–10004.
96. Sun, Y., Wang, J., Li, W., Zhang, J., Zhang, Y., & Fu, Y. (2015). DNA-stabilized bimetallic
nanozyme and its application on colorimetric assay of biothiols. Biosensors &
Bioelectronics, 74, 1038–1046.
97. Zhao, Q., Huang, H., Zhang, L., Wang, L., Zeng, Y., Xia, X., et al. (2016). Strategy to
fabricate naked-eye readout ultrasensitive plasmonic nanosensor based on enzyme mimetic
gold nanoclusters. Analytical Chemistry, 88, 1412–1418.
98. Tao, Y., Lin, Y. H., Huang, Z. Z., Ren, J. S., & Qu, X. G. (2013). Incorporating graphene
oxide and gold nanoclusters: A synergistic catalyst with surprisingly high peroxidase-like
activity over a broad ph range and its application for cancer cell detection. Advanced
Materials, 25, 2594–2599.
99. Song, Y. J., Xia, X. F., Wu, X. F., Wang, P., & Qin, L. D. (2014). Integration of platinum
nanoparticles with a volumetric bar-chart chip for biomarker assays. Angewandte
Chemie-International Edition, 53, 12451–12455.
100. Song, Y., Wang, Y., & Qin, L. (2013). A multistage volumetric bar chart chip for visualized
quantification of DNA. Journal of the American Chemical Society, 135, 16785–16788.
101. Gao, L., Liu, M. Q., Ma, G. F., Wang, Y. L., Zhao, L. N., Yuan, Q., et al. (2015).
Peptide-conjugated gold nanoprobe: Intrinsic nanozyme-linked immunsorbant assay of
integrin expression level on cell membrane. ACS Nano, 9, 10979–10990.
102. Han, M., Liu, S., Bao, J., & Dai, Z. (2012). Pd nanoparticle assemblies-as the substitute of
HRP, in their biosensing applications for H2O2 and glucose. Biosensors & Bioelectronics, 31,
151–156.
103. Jiang, X., Sun, C. J., Guo, Y., Nie, G. J., & Xu, L. (2015). Peroxidase-like activity of
apoferritin paired gold clusters for glucose detection. Biosensors & Bioelectronics, 64,
165–170.
104. Wang, Q., Zhang, L., Shang, C., Zhang, Z., & Dong, S. (2016). Triple-enzyme mimetic
activity of nickel-palladium hollow nanoparticles and their application in colorimetric
biosensing of glucose. Chemical Communications, 52, 5410–5413.
105. Lan, J., Xu, W., Wan, Q., Zhang, X., Lin, J., Chen, J., et al. (2014). Colorimetric
determination of sarcosine in urine samples of prostatic carcinoma by mimic enzyme
palladium nanoparticles. Analytica Chimica Acta, 825, 63–68.
106. Zheng, X. X., Liu, Q., Jing, C., Li, Y., Li, D., Luo, W. J., et al. (2011). Catalytic gold
nanoparticles for nanoplasmonic detection of DNA hybridization. Angewandte
Chemie-International Edition, 50, 11994–11998.
107. Luo, W. J., Zhu, C. F., Su, S., Li, D., He, Y., Huang, Q., et al. (2010). Self-catalyzed,
self-limiting growth of glucose oxidase-mimicking gold nanoparticles. ACS Nano, 4,
7451–7458.
108. Sharma, T. K., Ramanathan, R., Weerathunge, P., Mohammadtaheri, M., Daima, H. K.,
Shukla, R., et al. (2014). Aptamer-mediated ‘turn-off/turn-on’ nanozyme activity of gold
nanoparticles for kanamycin detection. Chemical Communications, 50, 15856–15859.
109. Weerathunge, P., Ramanathan, R., Shukla, R., Sharma, T. K., & Bansal, V. (2014).
Aptamer-controlled reversible inhibition of gold nanozyme activity for pesticide sensing.
Analytical Chemistry, 86, 11937–11941.
110. Daniel, M. C., & Astruc, D. (2004). Gold nanoparticles: Assembly, supramolecular
chemistry, quantum-size-related properties, and applications toward biology, catalysis, and
nanotechnology. Chemical Reviews, 104, 293–346.
111. Zhou, P., Jia, S., Pan, D., Wang, L., Gao, J., Lu, J., et al. (2015). Reversible regulation of
catalytic activity of gold nanoparticles with DNA nanomachines. Scientific Reports, 5,
14402.
112. Hizir, M. S., Top, M., Balcioglu, M., Rana, M., Robertson, N. M., Shen, F. S., et al. (2016).
Multiplexed activity of perauxidase: DNA-capped aunps act as adjustable peroxidase.
Analytical Chemistry, 88, 600–605.
References 55

113. Lien, C. W., Huang, C. C., & Chang, H. T. (2012). Peroxidase-mimic bismuth-gold
nanoparticles for determining the activity of thrombin and drug screening. Chemical
Communications, 48, 7952–7954.
114. Zhan, P., Wang, J., Wang, Z.-G., & Ding, B. (2014). Engineering the pH-responsive catalytic
behavior of aunps by DNA. Small (Weinheim an der Bergstrasse, Germany), 10, 399–406.
115. Hu, J. T., Ni, P. J., Dai, H. C., Sun, Y. J., Wang, Y. L., Jiang, S., et al. (2015). A facile
label-free colorimetric aptasensor for ricin based on the peroxidase-like activity of gold
nanoparticles. RSC Advances, 5, 16036–16041.
116. Ni, P. J., Dai, H. C., Wang, Y. L., Sun, Y. J., Shi, Y., Hu, J. T., et al. (2014). Visual detection
of melamine based on the peroxidase-like activity enhancement of bare gold nanoparticles.
Biosensors & Bioelectronics, 60, 286–291.
117. Li, W., Bin, C., Zhang, H. X., Sun, Y. H., Wang, J., Zhang, J. L., et al. (2015).
BSA-stabilized Pt nanozyme for peroxidase mimetics and its application on colorimetric
detection of mercury(II) ions. Biosensors & Bioelectronics, 66, 251–258.
118. Fu, Y., Zhang, H. X., Dai, S. D., Zhi, X., Zhang, J. L., & Li, W. (2015).
Glutathione-stabilized palladium nanozyme for colorimetric assay of silver(I) ions.
Analyst, 140, 6676–6683.
119. Tao, Y., Ju, E. G., Ren, J. S., & Qu, X. G. (2015). Bifunctionalized mesoporous
silica-supported gold nanoparticles: Intrinsic oxidase and peroxidase catalytic activities for
antibacterial applications. Advanced Materials, 27, 1097–1104.
Chapter 4
Metal Oxide-Based Nanomaterials
for Nanozymes

Abstract Metal oxide-based nanomaterials have been extensively studied to mimic


various natural enzymes due to their unique properties. In this chapter, several metal
oxide-based nanozymes are discussed. First, the use of cerium oxide nanomaterials
for mimicking natural enzymes (such as superoxide dismutase, catalase, oxidase,
peroxidase, phosphatase, etc.) is discussed. Second, the use of iron oxide nano-
materials for peroxidase mimics and other mimics is covered. Third, the enzyme
mimicking activities of other metal oxides (such as vanadium oxide, cobalt oxide,
copper oxide, etc.) are discussed. The catalytic mechanisms are also discussed if
they have been elucidated. Selected examples for broad applications are discussed,
which cover from glucose detection, DNA detection, immunoassay, and
immunostaining, to neuroprotection, cardioprotection, cancer therapy, and tissue
engineering.

Keywords Nanozymes 
Artificial enzymes 
Enzyme mimics Metal 
  
oxide-based nanomaterials Cerium oxide Iron oxide Reactive oxygen species 
 
Oxidase mimics Glucose detection Brain chemistry

Metal oxide-based nanomaterials (such as cerium oxide and iron oxide) have been
extensively studied to mimic various natural enzymes. These nanozymes have
found broad applications in many areas, from bioanalysis to therapeutics. In this
chapter, we will discuss the nanozymes based on these metal oxide nanomaterials.

4.1 Cerium Oxide

Cerium oxide nanomaterials (also called nanoceria) have been widely used as
highly efficient catalysts due to their unique properties (i.e., the presence of mixed
valence states of Ce3+ and Ce4+ and highly mobile lattice oxygen) [1]. They have
also been extensively explored for mimicking natural enzymes [2–8]. In 2005,
Tarnuzzer, Seal and co-workers reported that vacancy engineered nanoceria could

© The Author(s) 2016 57


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_4
58 4 Metal Oxide-Based Nanomaterials for Nanozymes

protect normal but not tumor cells from radiation-induced damage [9]. Since then, a
large number of studies have showed that nanoceria could mimic SOD, catalase,
oxidase, peroxidase, phosphatase, etc. [10–35].

4.1.1 Cerium Oxide as SOD Mimics

(a) Catalytic activities and mechanisms


In their seminal study, Tarnuzzer, Seal and co-workers attributed the protective
role of the nanoceria to its free radicals elimination capability. It was suggested that
the radicals could be eliminated by nanoceria via a Ce3 þ ! Ce4 þ ! Ce3 þ
regeneration mechanism [9]. Self, Seal, and co-workers as well as other groups
have then established the SOD mimicking activities of nanoceria [36–38]. Based on
the competitive cytochrome C assay, H2O2 formation assay, and EPR measure-
ments, Self et al. [36, 37] proposed that the nanoceria would exert the SOD-like
activities via the mechanisms shown in Eqs. 4.1–4.2. A possible molecular
mechanism was proposed (Fig. 4.1) [39]. However, more detailed studies are still
needed to confirm (or even modify) the proposed mechanism.

Fig. 4.1 Proposed molecular mechanisms for the SOD mimicking activity of nanoceria.
Reprinted from Ref. [39], Copyright 2011, with permission from Royal Society of Chemistry
4.1 Cerium Oxide 59

O
2 þ Ce

! O2 þ Ce3 þ ð4:1Þ

O
2 þ Ce

þ 2H þ ! H2 O2 þ Ce4 þ ð4:2Þ

The SOD-like catalytic activity of nanoceria was dependent on the redox state of
surface cerium ions (i.e., the SOD-like activity was positively correlated with the
ratio of Ce3+/Ce4+) [36, 37]. It was also found that the SOD mimicking activity of
nanoceria was size dependent. Nanoceria larger than 5 nm did not show prominent
SOD activity [36]. An interesting strategy has been proposed to endow the
nanoceria larger than 5 nm with SOD mimicking activities [40]. When electrons
were transferred either from a native SOD or other donors (e.g., an inorganic metal
complex) to the nanoceria with large size, its superoxide-scavenging capability was
restored. The transferred electrons would reduce Ce4+ to Ce3+ and thus improve the
SOD-like activity [40]. The SOD mimicking activities of nanoceria could be
modulated by other strategies [41–43]. For example, by doping with nanoceria with
redox inactive Sm and Ti atoms, its catalytic activity could be reduced [41, 43].
(b) Applications
Nanoceria-based SOD mimics have been used for various applications, ranging
from anti-inflammation and neuroprotection to tissue engineering and cancer
therapy [2, 44–46].
Neuroprotection. Numerous studies have demonstrated that nanoceria had
neuroprotection functions [47–50]. In their pioneering study, McGinnis et al.
showed that nanoceria could protect rat retina photoreceptor cells from light-induced
degeneration (Fig. 4.2). The protection activity was probably due to the elimination
of ROS (reactive oxygen species) by nanoceria. Encouragingly, the nanozymes
showed therapeutic effects when they were administrated either before or even after
light exposure [47]. In a later report, they showed that nanoceria exhibited long-term
therapeutic effects of protecting photoreceptor cells from degeneration [48]. Using
the very low density lipoprotein receptor knockout (vldlr–/–) mouse as a model,
McGinnis et al. [49] demonstrated that nanoceria was effective against age-related
macular degeneration diseases.
Other groups have also demonstrated the neuroprotection roles of nanoceria in
several disease models, such as adult rat spinal cord injury, brain ischemia,
Alzheimer’s disease, autoimmune encephalomyelitis, etc. [51–55]. For example,
using mouse hippocampal brain slice as an in vitro model of brain ischemia, it has
showed that the treatment of nanoceria could significantly reduce the ischemic cell
death (Fig. 4.3) [52]. Detailed studies revealed that the concentrations of super-
oxide (O∙−
2 ) and nitric oxide were reduced by around 15 %. Moreover, the level of
peroxynitrite-induced 3-nitrotyrosine was remarkably reduced by 70 %. Therefore,
it was proposed that the nanozymes may protect the brain cells from ischemic injury
by scavenging ROS and NOS (such as peroxynitrite, O∙− 2 , and nitric oxide) [52].
The nanozyme even protected cell from ischemic injury in living mice brains [53].
60 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.2 Intravitreal injection of nanoceria protected rat retina photoreceptor cells from
light-induced degeneration. Reprinted from Ref. [47], Copyright 2006, with permission from
Nature Publishing Group

To further improve the therapeutic efficacy, one may conjugate specific recognition
moieties onto the nanozymes for targeted therapy [55].
Cardioprotection. A few studies have showed that nanoceria possessed car-
dioprotective activities [56–58]. Traversa, Nardo, and co-workers showed that
nanoceria protected cardiac progenitor cells, a valuable cell source for cardiac
regenerative medicine, from oxidative stress [56]. For an in vivo mouse model with
ischemic cardiomyopathy, the treatment with nanoceria effectively inhibited the
progressive left ventricular dysfunction and dilatation. The therapeutic benefits of
the nanozymes were due to the inhibition of myocardial oxidative stress, ER stress
and inflammatory processes, which were evidenced by the decrease of
pro-inflammatory cytokines (such as tumor necrosis factor-α and interleukin-1β)
and downregulation of endoplasmic reticulum stress-associated genes [59]. Heart
hypertrophy following pulmonary arterial hypertension could also be attenuated by
nanoceria treatment [58].
Cancer therapy. Researchers have used nanoceria to treat cancers [60–63].
Cellular studies demonstrated the feasibility of using nanoceria for cancer therapy.
Due to the Warburg effect, the nanozyme would benefit stromal cells by eliminating
the elevated ROS. At the same time, the nanozyme would inhibit tumor cells’
metastatic spread by inhibiting the myofibroblasts formation, etc. [60]. The in vivo
anticancer activities of the nanoceria were evaluated using mice with xenografted
melanoma. The treatment with nanoceria led to significantly smaller tumor volume
4.1 Cerium Oxide 61

Fig. 4.3 Neuroprotective effect of nanoceria treatment after brain ischemia. Nanoceria signifi-
cantly decreased the area of ischemia-induced cell death in hippocampal slices. a Pseudocolor
images of brain slices loaded with Sytox blue. b Quantification of the area of Sytox fluorescence in
mouse hippocampal slices after 30 min of ischemia and treatment with varying doses of nanoceria.
Reprinted from Ref. [52], Copyright 2011, with permission from Elsevier

and weight, thus showing the translational promise of the nanoceria [61]. It was also
reported that nanoceria could inhibit ovarian tumor growth via an anti-angiogenic
mechanism [62].
Tissue engineering. Nanoceria has been used to promote (stem) cell prolifera-
tion and tissue engineering [64–68]. Traversa et al. prepared hybrid materials with
enhanced mechanical properties by fabricating nanoceria together with PLGA
scaffolds. They went on to demonstrate that the hybrids could promote the
murine-derived cardiac and mesenchymal stem cells growth. Initially, the promo-
tion effects of nanoceria were attributed to its antioxidation properties [64].
However, later they revealed that the cell proliferation was more favored on Ce4+
dominant regions rather than on Ce3+ dominant regions [65]. It suggested that Ce4+
dominant regions would promote the cell proliferation due to its hydrophilicity. On
the other hand, Ce3+ dominant regions inhibited the cell proliferation by limiting
62 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.4 Specific cell growth process on Ce4+and Ce3+ regions. Reprinted from Ref. [65],
Copyright 2014, with permission from Elsevier

cell spreading and weakening the cell–material interactions (Fig. 4.4) [65]. This
result was in agreement with other reports, which suggested that surface Ce3+ might
be biotoxic [69].
Mattson, Seal, and co-workers demonstrated that nanoceria could accelerate
cutaneous wound healing in mice by promoting the growth of keratinocytes,
fibroblasts, and vascular endothelial cells. The therapeutic efficacy was owing to the
enhanced proliferation and migration of these cells [66]. As shown in Fig. 4.5,
when nanoceria (prepared in water) was used as additives, they could promote bone
regeneration by enhancing collagen production from human mesenchymal stem
cells. In comparison with the scaffolds themselves, osteoblast-like products were
observed on the nanoceria-doped scaffolds even in the absence of osteogenic
supplements. The promotion mechanism still remained unclear [67].
Anti-inflammation and antioxidation. Owing to the mixed valence and oxygen
vacancy, nanoceria-based nanozymes exhibited anti-inflammation and antioxida-
tion effects. Using J774A.1 murine macrophage cells, Reilly et al. demonstrated
that nanoceria possessed anti-inflammatory activities. The nanoceria exhibited good
biocompatibility after being internalized by the cells. Moreover, the nanoceria could
both scavenge ROS and inhibit NO production, which in turn would inhibit
inflammation [70]. In a mouse model with CCl4-induced liver injury, it was showed
that the treatment with nanoceria mitigated systemic inflammation, reduced hepatic
steatosis, and improved portal pressure. The anti-inflammatory effects of the
nanoceria were also evidenced by the remarkable downregulation in mRNA
expression of inflammatory cytokines, vasoconstrictor, oxidative stress messengers,
etc. This study indicated that nanoceria might be used as a therapeutic drug for liver
disease [71]. Nanoceria has also been widely used as antioxidants [41, 56, 72, 73].
Its antioxidizing activity was even better than that of commercial antioxidants such
as Trolox [73].
Other applications. Nanoceria-based SOD mimics have been explored for other
interesting applications [74, 75]. For instance, nanoceria showed antibacterial
activities against Escherichia coli (a typical Gram-negative bacterium) and Bacillus
subtilis (a typical Gram-positive bacterium) [74]. More interestingly, it showed that
the nanoceria was ineffective against Shewanella oneidensis, which was not only a
Gram-negative bacterium but also a metal-reducing bacterium. These studies
indicated that the antibacterial activities of nanoceria were dependent not only on
4.1 Cerium Oxide 63

Fig. 4.5 SEM images of human mesenchymal stem cells cultured for 10 days on the bioactive
glass scaffolds doped with nanoceria synthesized in water (a, b), in dextran (c, d) and without
nanoceria (e, f). Cell culture was performed in the absence (a, c, e) and in the presence (b, d, f) of
osteogenic supplements. Cells attached and spread on the scaffolds’ surface. Note the
osteoblast-like products or minerals on the cells cultured in the absence (a, c, e) of osteogenic
supplements (scale bar = 10 μm). Reprinted from Ref. [67], Copyright 2010, with permission
from Royal Society of Chemistry

the nanozyme’s intrinsic properties (such as size, surface coating, etc.) but also on
the bacterial species.
Recently, nanoceria was integrated with a bioresorbable electronic stent [76].
During percutaneous coronary interventions, it might produce ROS and cause
in-stent thrombosis due to the inflammation. To address these issues, nanoceria was
used to not only scavenge ROS but also mitigate inflammation.
64 4 Metal Oxide-Based Nanomaterials for Nanozymes

4.1.2 Cerium Oxide as Catalase Mimics

Besides the SOD mimicking activities, the catalase-like activities of nanoceria have
also been studied [77–80]. It was found that the oxidase mimicking activity of
nanoceria was negatively correlated with the ratio of Ce3+/Ce4+, which was different
from its SOD mimicking activity [78]. Detailed experimental and computational
studies revealed that nanoceria could dismutate H2O2 into H2O and O2 via the
mechanism shown in Eqs. 4.3–4.7 [79]. Such a mechanism was also consistent with
a previously proposed one (Fig. 4.6) [39].

2Ce3 þ þ H2 O2 þ 2H þ ! 2H2 O þ 2Ce4 þ ð4:3Þ

H2 O2 þ Ce4 þ ! Ce3 þ þ 2H þ þ HOO ð4:4Þ

Ce4 þ þ HOO ! Ce3 þ þ H þ þ O2 ð4:5Þ

H2 O2 þ 2Ce4 þ ! 2Ce3 þ þ 2H þ þ O2 ð4:6Þ

2H2 O2 ! 2H2 O þ O2 ð4:7Þ

Fig. 4.6 Proposed molecular mechanisms for the catalase mimicking activity of nanoceria.
Reprinted from Ref. [39], Copyright 2011, with permission from Royal Society of Chemistry
4.1 Cerium Oxide 65

Traditional bioactive glasses with the composition of Na2O, CaO, SiO2, and
P2O5 have been used for bone defect reparation. However, they cannot effectively
prevent oxidative stress and reduce inflammation after implantation due to the inert
chemistry. By introducing redox active nanomaterials, especially the ones with
antioxidation and anti-inflammation activities, into the bioactive glasses would
address this issue. To this end, nanoceria has been used to dope traditional bioactive
glasses. The catalase mimicking activity of nanoceria-doped glasses was confirmed
by in vitro experiments and computational studies [79].
Liu et al. [77] developed an interesting strategy to fabricate biosensors for H2O2
and glucose. As shown in Fig. 4.7, the adsorption of dye-labeled DNA onto
nanoceria quenched its fluorescence. The presence of H2O2 would displace

Fig. 4.7 a Sensing H2O2 by displacing adsorbed fluorescent DNA from nanoceria. The color of
nanoceria is changed in the same process. b A proposed mechanism of H2O2-induced DNA release
by capping the nanoceria surface. For the three time scales marked in the scheme, DNA release is
related to the one on the order of 1 min. Adapted from Ref. [77], Copyright 2015, with permission
from American Chemical Society
66 4 Metal Oxide-Based Nanomaterials for Nanozymes

adsorbed DNA from nanoceria and thus recover its fluorescence. The possible
sensing mechanism was showed in Fig. 4.7b. The phosphate group of the DNA
interacted with the Ce3+ on the nanoceria surface. H2O2, however, could compete
against the bound DNA by oxidizing the Ce3+ into Ce4+ within one minute. The
bound H2O2 would then be converted into H2O and O2 via a catalase-like mech-
anism. Moreover, when glucose oxidase (GOx) was coupled with the H2O2-
mediated DNA displacement from nanoceria, sensitive and selective glucose
detection was achieved [77]. Using the same strategy, cellular H2O2 was detected
with cerium oxide nanowires [80].

4.1.3 Cerium Oxide as Peroxidase Mimics

Nanoceria has been successfully used to mimic peroxidase [5, 81, 82]. The per-
oxidase mimicking activities were confirmed by the fact that nanoceria could cat-
alyze peroxidase’s substrates with H2O2. The detection of H2O2 and glucose has
been achieved by making use of the nanoceria’s peroxidase-like activities [5].
Moreover, a sandwich immunoassay has been developed for detection of a breast
cancer biomarker (i.e., CA15-3) [81]. The peroxidase-like nanoceria was attached
to the detection antibody electrostatically, eliminating the time-consuming covalent
bioconjugation procedure. Compared with the natural HRP-based sandwich assay,
the nanozyme-based assay not only exhibited higher stability and lower cost but
also had higher sensitivity (i.e., it had a detection limit one order of magnitude
lower than that of HRP-based assay) [81].

4.1.4 Cerium Oxide as Oxidase Mimics

Perez et al. [83] reported the oxidase mimicking activity of polymer-coated


nanoceria. Since no H2O2 was involved, the observed catalytic activity toward
TMB oxidation at acidic conditions was attributed to the oxidase-like rather
peroxidase-like properties of the nanoceria. The size-dependent study showed that
smaller nanoceria exhibited higher oxidase mimicking activities. Moreover, the
catalytic activities were also dependent on the surface coatings. The thinner and
more permeable poly(acrylic acid)-coated nanoceria exhibited higher activity in
comparison with the nanoceria coated with thicker dextran. Interestingly, when
nanoceria was conjugated with folate, the obtained nanozyme probes were used for
specific cancer cell detection (Fig. 4.8) [83]. The developed immunoassay was
advantageous over traditional ELISA in the eliminating the use of unstable H2O2
and easily denatured HRP. The oxidase mimicking activities of nanoceria could be
regulated by binding with ssDNA [84]. A colorimetric method for DNA detection
has been reported by using the oxidase-like nanoceria. The detection was based on
the target DNA induced shielding of the catalytic activity of nanozyme [85].
4.1 Cerium Oxide 67

Fig. 4.8 Comparison of traditional ELISA (a) and nanoceria-based ELISA (b). In traditional
ELISA, an HRP antibody is utilized as secondary antibody that, upon hydrogen peroxide
treatment, facilitates the oxidation of TMB, resulting in color development. In nanoceria-based
ELISA, the oxidase-like activity of nanoceria facilitates the direct oxidation of TMB without the
need of HRP or hydrogen peroxide. Reprinted from Ref. [83], Copyright 2009, with permission
from John Wiley and Sons

4.1.5 Cerium Oxide as Other Mimics

Nanoceria has also been explored for mimicking other enzymes [18, 86, 87]. For
example, nanoceria was able to eliminate nitric oxide radical. Unexpectedly, the
nitric oxide radical scavenging activity of nanoceria was negatively correlated with
the Ce3+/Ce4+ ratio, which was different from its SOD-like activity [86].
Kuchma et al. found that the nanoceria exhibited phosphatase-like activities.
Nanoceria promoted the hydrolysis of phosphate ester bonds in p-nitrophenyl-
phosphate (pNPP), o-phospho-L-tyrosine, ATP, but not in DNA. They showed that
the phosphatase mimicking activity was dependent on the Ce3+ sites and oxygen
vacancies. Facilitated with first principles calculation, it was revealed that the
catalysis proceeded via a SN2 mechanism (Fig. 4.9). Moreover, the activation
energy for the hydrolysis was significantly reduced when it was mediated by a Ce3
+
-Ce3+ complex instead of a Ce4+-Ce4+ complex (from 22.9 to 13.6 kcal/mol). The
lowered activation energy was attributed to the less polarized phosphorus-oxygen
bonds in the phosphate-Ce3+-Ce3+ complex, which was purely electronic. It also
suggested that the disfavored DNA hydrolysis with the nanoceria might be due to
the steric hindrance, which could shield the interaction of phosphate bond with the
Ce3+ sites [18].
68 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.9 The snapshots of the reaction pathway for pNPP hydrolysis catalyzed by nanoceria.
a The reaction complex includes the neutral water molecule in the secondary coordination sphere.
b The hydrogen bonding is switched to the bridging hydroxide anion. c The proton is transferred
along this newly formed hydrogen bond. d The formed uncoordinated hydroxide then attacks
pNPP and e forms the transition state. f The p-nitrophenyl anion then leaves, g deprotonates one of
the water molecules in the first coordination sphere, and h forms the hydrogen-bonded product
complex. Reprinted from Ref. [18], Copyright 2010, with permission from Elsevier

4.2 Iron Oxide

Owing to their superior magnetic properties, iron oxide nanomaterials have been
widely used in bioanalytical and biomedical fields for separation and capture of
analytes, etc. Before Yan and co-workers’ surprising discovery, it was believed that
these nanomaterials were chemically and biologically inert. Therefore, they were
usually conjugated with various functional groups (such as metal catalysts,
enzymes, or antibodies) for practical applications. In 2007, Yan et al. discovered the
unexpected peroxidase mimicking activity of Fe3O4 magnetic nanoparticles
(MNPs) [88]. Inspired by this seminal work, quite a lot of nanomaterials (including
the iron oxide nanomaterials) have been explored to study their enzyme mimicking
activities [6, 7, 89–115].
Most of iron oxide nanomaterials have showed intrinsic peroxidase mimicking
activities though they could also mimic other natural enzymes. In this section, the
enzyme mimetic properties of iron oxide nanomaterials and their applications are
covered.

4.2.1 Iron Oxide as Peroxidase Mimics

(a) Catalytic activities and mechanisms


Natural peroxidases catalyze the oxidation of its substrates with peroxide (H2O2
in most of case). They play important roles in biological systems. For example,
4.2 Iron Oxide 69

Fig. 4.10 Fe3O4 MNPs performed as peroxidase mimics. a TEM images of employed Fe3O4
MNPs with different sizes. b The colorless peroxidase substrates TMB, DA and OPD were oxidized
into their corresponding colored products in the presence of H2O2 via the catalysis of Fe3O4 MNPs.
Reprinted from Ref. [88], Copyright 2007, with permission from Nature Publishing Group

glutathione (GSH) peroxidase can detoxify ROS while myeloperoxidase can defend
against pathogens. Moreover, HRP is widely used in bioanalytical and biomedical
research for amplifying detection signals, where it is conjugated to an antibody or
other biorecognition molecules. In their pioneering study, Yan et al. showed that
Fe3O4 MNPs could catalyze the oxidation of several peroxidase substrates with
H2O2, suggesting that they possessed intrinsic peroxidase mimicking activities
(Fig. 4.10) [88]. Compared with natural HRP, the nanozymes were unique in
several aspects. First, they were more stable and could work in wider pH and
temperature conditions. Second, they could be produced in large scale with low
cost. Third, they could be conjugated with biorecognition (or other functional)
elements due to the large surface area and rich surface chemistry. Fourth, they were
multifunctional (i.e., they had both magnetic and catalytic properties) [88].
Fe3O4 MNPs with different sizes (30, 50 and 300 nm) were studied, showing
that smaller nanoparticles had higher catalytic activity. Enzyme kinetics study
revealed that the Fe3O4 MNPs mimicked reactions also followed a ping-pong
mechanism (i.e., they exhibited parallel double-reciprocal plots of substrate con-
centration versus reaction rate) [88]. Interestingly, the nanozyme and HRP had Km
values of 0.098 mM versus 0.434 mM for TMB, respectively, which indicated that
the nanozyme showed even higher affinity to TMB in comparison with HRP. On
the other hand, the nanozyme exhibited lower affinity to H2O2 compared with HRP
[88]. The exact molecular mechanisms for the Fe3O4 MNPs mimicking peroxidase
activities are still unclear. A few studies have suggested that Fenton and/or Haber–
Weiss reaction mechanisms could be involved [116–118].
70 4 Metal Oxide-Based Nanomaterials for Nanozymes

(b) Applications
Iron oxide nanomaterial-based nanozymes have been applied for many inter-
esting applications, ranging from detection of H2O2 and glucose to immunoassay
and immunostaining.
H2O2 detection. Detection of H2O2 is very important because it plays critical
roles in biology, medicine, environmental protection, food industry, etc. Wei and
Wang have reported the first example of H2O2 detection using Fe3O4 MNPs as the
peroxidase mimic and ABTS as the colorimetric substrate (Fig. 4.11a) [119, 120].
The presence of H2O2 could be easily visualized by the colored product of ABTSox
(i.e., oxidized ABTS∙+). Moreover, the amount of H2O2 could be quantified by
measuring the corresponding absorption spectra [119]. Since then, considerable
studies have been devoted to the H2O2 detection using peroxidase-like nanozymes
(see Table A.1) [6, 7].

Fig. 4.11 a Nanozyme as peroxidase mimic for colorimetric sensing of H2O2, and glucose when
combined with glucose oxidase. b The sensing format in (a) could be extended to other targets
(substrate 1 here) when combined with a suitable oxidase. c Target of interest as substrate 0 could
be determined if it could be converted into an oxidase substrate. Numerous transduction signals
can be adopted for sensing (such as colorimetric, fluorometric, chemiluminescent, and SERS
signals when the corresponding substrates are used; and electrochemical signals when a nanozyme
is immobilized on an electrode). Reprinted from Ref. [7], Copyright 2016, with permission from
Royal Society of Chemistry
4.2 Iron Oxide 71

Glucose (and other oxidase substrate) detection. Wei and Wang [119] went on
to demonstrate that glucose detection could be achieved by coupling GOx with a
peroxidase-like nanozyme. As shown in Fig. 4.11a, GOx first converted glucose to
H2O2, which subsequently oxidized a substrate into the corresponding product
(such as colored ABTS∙+ here). The established method showed excellent sensi-
tivity and selectivity toward glucose detection (Fig. 4.12) [119]. Followed by this
work, others have used other nanomaterials-based peroxidase mimics to detect
glucose even in complicated samples (such as in serum, urine, drinks, etc.) (see
Table A.2 for more examples) [6, 7].
As indicated in Fig. 4.11b, when other oxidases were used, the corresponding
oxidase substrates could be detected [6, 7]. Numerous biological important small
molecules, such as choline, uric acid, D-alanine, lactate, and xanthine, have been
detected with the proposed sensing method (see Table A.2 for more examples) [6, 7].
As indicated in Fig. 4.11c, the analytes as substrate 0 could also be detected by
converting them to oxidase substrates [6, 7]. Moreover, the proposed sensing
strategy could also be used to screen enzyme inhibitors. For example, Yan et al.
[121] reported a rapid and sensitive method for detecting organophosphorus pes-
ticides and nerve agents (such as Sarin), which were high potent acetyl-
cholinesterase inhibitors.
DNA detection. Two possible approaches are available for DNA detection by
using the peroxidase mimicking activities of iron oxide nanomaterials. First, it has
been established that ssDNA and dsDNA exhibited different affinities toward the

Fig. 4.12 Colorimetric detection of glucose by combining GOx with Fe3O4 MNPs as a
peroxidase mimic. Reprinted from Ref. [119], Copyright 2008, with permission from American
Chemical Society
72 4 Metal Oxide-Based Nanomaterials for Nanozymes

nanozymes. A probe ssDNA would adsorb onto a nanozyme and thus shield its
activity. The presence of a target ssDNA would form a duplex with the probe
ssDNA, which would subsequently recover the nanozyme’s activity. For instance,
Park et al. reported the detection of Chlamydia trachomatis pathogen DNA from a
human urine sample via such a sensing strategy [122]. Second, by using a nano-
zyme as an alternative to conventional dyes (or enzymes) to label a probe ssDNA,
the target DNA could be detected via by hybridizing the probe ssDNA and target
ssDNA. Many sensing platforms could be adopted for this purpose. For example, in
a sandwich assay, a capture ssDNA would first interact with a target ssDNA. Then a
nanozyme labeled probe ssDNA would further interact with the captured target
ssDNA for signaling [123].
Liu et al. [124] found that in certain cases, DNA could even accelerate the
peroxidase mimicking activities of Fe3O4 MNPs. It was speculated that the
enhancement might be due to the electrostatic interaction between negatively
charged DNA and positively charged TMB substrate.
Immunoassay. In Yan’s initial report, they developed two immunoassay formats
[88]. Using an antigen-down immunoassay format, the detection of hepatitis B virus
surface antigen (preS1) was achieved. In a second demonstration, a capture-
detection sandwich immunoassay format was used for the detection of the
myocardial infarction biomarker troponin I (TnI) [88]. Yan and co-workers recently
developed a nanozyme-strip for Ebola detection by integrating lateral flow tech-
nique with the Fe3O4 MNPs nanozyme (Fig. 4.13) [125]. AuNPs are commonly
used to fabricate lateral flow strips, but their sensitivity should be further improved
to meet the requirements for infectious diseases (such as Ebola) diagnosis and
monitoring. To tackle this challenge, Yan et al. used Fe3O4 MNPs nanozyme to
replace AuNPs for fabricating the nanozyme-strips. Compared with the
AuNPs-based strip, the nanozyme exhibited 100-fold better sensitivity toward the
detection of Ebola virus glycoprotein (EBOV-GP) (Fig. 4.13c–e). The enhanced
sensitivity could be attributed to the high catalytic activity of the nanozymes.
Moreover, the nanozyme-strip showed comparable analytical performance with
ELISA. Due to high sensitivity, low fabrication cost, and ease of use, the
nanozyme-strip would find applications in fighting against emergent virus diseases,
especially in resource limited areas [125].
Other important targets, such as carcinoembryonic antigen (CEA), has been
detected by using Fe3O4 MNPs-based peroxidase mimics (see Table A.3 for more
examples) [126]. Cancer cells have also been selectively detected when the nano-
zymes were labeled with specific biorecognition elements (such as folic acid) [127].
Immunostaining. Yan et al. [99] also developed nanozymes for specific tumor
tissue imaging (Fig. 4.14). The probe was prepared by encapsulating catalytic iron
oxide nanoparticles within recombinant human heavy-chain ferritin shells to get
magnetoferritin nanoparticles. The magnetoferritin nanoparticles exhibited peroxi-
dase mimicking activity. When they were used for staining tumor tissues, they
4.2 Iron Oxide 73

Fig. 4.13 a Standard AuNPs-based strip, b nanozyme-strip employing Fe3O4 MNPs in place of
AuNPs, c nanozyme-strip, d standard colloidal gold strip and e ELISA method for EBOV-GP
detection. The asterisk (*) indicates the limit of visual detection of the test line in strips.#
OD450 nm > cut-off value. Reprinted from Ref. [125], Copyright 2015, with permission from
Elsevier

would differentiate tumor tissues from the benign ones due to the specific inter-
actions between ferritin and the overexpressed transferrin receptor 1 onto tumor
tissues. Very encouragingly, for the 474 patient specimens stained, the magneto-
ferritin nanoparticles distinguished cancer samples from normal ones with a
remarkable success (i.e., with a sensitivity of 98 % and a specificity of 95 %). Gu
and co-workers also showed that iron oxide nanoparticles-based nanozymes could
be used for immunohistochemical studies [94].
Aptasensors. Several studies have demonstrated that aptasensors could be
developed by combining the specific recognition capability of aptamers and the
74 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.14 Magnetoferritin nanoparticles as peroxidase mimic for tumor tissue staining and
imaging. a Preparation of magnetoferritin nanoparticles. b Magnetoferritin nanoparticle staining of
tumor tissues. Reprinted from Ref. [99], Copyright 2012, with permission from Nature Publishing
Group

high catalytic activities of iron oxide-based nanozymes [128–130]. For instance,


Yang et al. [128] reported an aptasensor for thrombin. It is well known that
thrombin has two specific binding aptamers, one is a 15-mer and the other is a
29-mer. The two aptamers bind to the different sites onto a thrombin and thus would
form a sandwich structure. Based on such specific interaction, the 29-mer aptamer
was used to capture thrombin while the 15-mer aptamer was used to label iron
oxide nanoparticles. The presence of thrombin would form a sandwich structure,
and the 15-mer aptamer labeled iron oxide nanoparticles would catalyze the oxi-
dation of TMB for signaling [128]. Zhu, Wang, and co-workers [129] developed a
peroxidase mimic based on hybrid nanostructures (Fig. 4.15a). The nanozyme had
a Fe3O4 core, coated with Ag–Pd nanocages. Electrochemical measurements
demonstrated that the hybrid nanostructures had higher catalytic activity than Fe3O4
alone. They then developed an aptasensor for CTCs (circulating tumor cells)
4.2 Iron Oxide 75

Fig. 4.15 a Schematic illustration of the fabrication of Fe3O4@Ag–Pd nanozymes. b Schematic


illustration of CTCs detection. c DPV responses to different concentrations of MCF-7 cells. d DPV
responses to various types of cells. Reprinted from Ref. [129], Copyright 2014, with permission
from American Chemical Society

detection. SYL3C aptamer, which specifically interact with overexpressed epithelial


cell adhesion molecule (EpCAM) on CTCs, was used to capture the CTCs and to
label the hybrid nanostructures (Fig. 4.15b). As shown in Fig. 4.15c, d, the
developed aptasensor showed good sensitivity and selectivity toward CTCs
detection [129].
76 4 Metal Oxide-Based Nanomaterials for Nanozymes

4.2.2 Iron Oxide as Other Enzyme Mimics

Iron oxide nanoparticles have been used to mimic catalase as well as other enzymes
[108, 131–135]. For instance, Gu and co-workers [131] found that iron oxide
nanoparticles exhibited dual enzyme-like activities. For both Fe3O4 and γ-Fe2O3
nanoparticles, they showed peroxidase-like and catalase-like activities at acidic and
neutral conditions, respectively. It was found that both the peroxidase and catalase
mimicking activities of Fe3O4 nanoparticles were higher than that of γ-Fe2O3
nanoparticles. The cellular studies revealed that the iron oxide nanoparticles
exhibited concentration-dependent cytotoxicity to human glioma U251 cells. The
cytotoxicity was attributed to the nanoparticles’ peroxidase-like activities within
acidic lysosomes, which would trap the nanoparticles to catalytically produce
hydroxyl radicals (Fig. 4.16). It also suggested that γ-Fe2O3 would be safer for
biomedical applications [131].
In a subsequent study, Gu and co-workers [133] went on to use Fe2O3
nanoparticles to protect hearts from ischemic damage both in vitro and in vivo
(Fig. 4.17a). It revealed that the protective activity was mainly from the nanopar-
ticles themselves rather than the surface coatings. Ischemic injury is mediated via
complicated mechanisms. Therefore, comprehensive studies were carried out,
which suggested that the Fe2O3 nanoparticles would exert their protection effects
via several mechanisms. First, they could inhibit the cellular ROS-induced mem-
brane lipid peroxidation. Second, they could attenuate Ca2+ influx, which in turn

Fig. 4.16 Schematic illustration of peroxidase-like activity-induced cytotoxicity by iron oxide


nanoparticles. The nanoparticles are trapped in acidic lysosomes when internalized into cells, so
they catalyze H2O2 to produce hydroxyl radicals through peroxidase-like activity; however, in
neutral cytosol, the nanoparticles would decompose H2O2 through catalase-like activity. Reprinted
from Ref. [131], Copyright 2012, with permission from American Chemical Society
4.2 Iron Oxide 77

Fig. 4.17 a Cardioprotective activity of Fe2O3 nanoparticles. b Effects of dietary Fe3O4


nanoparticles in a Drosophila Alzheimer’s Disease model. a Reprinted from Ref. [133], Copyright
2015, with permission from Nature Publishing Group. b Reprinted from Ref. [132], Copyright
2016, with permission from John Wiley and Sons

inhibited ROS. Third, they could increase NO production by promoting NO pro-


ducing proteins’ activities and enhance NO protective effects by increasing the level
of S-nitrosothiols. More convincing investigations are needed to completely
understand the protective roles of the Fe2O3 nanoparticles [108]. Nevertheless,
since the nanoparticles did not show obvious toxicity toward normal cardiomy-
ocytes and they were more potent than Verapamil (a synthetic drug) and Salvia
miltiorrhiza extract (a natural antioxidant), they were expected to be further
explored as potential nanomedicine for cardiovascular diseases treatment [133]. The
protective effects of iron oxide nanoparticles against ischemic brain injury were also
reported [134].
Dietary Fe3O4 nanoparticles have been used to treat Drosophila with Alzheimer’s
Disease (Fig. 4.17b) [132]. It was demonstrated that the Fe3O4 nanoparticles could
mimic catalase in vivo. Therefore, they would improve neurodegeneration in a
Drosophila Alzheimer’s Disease model by reducing intracellular oxidative stress
[132]. This study together with others indicated that iron oxide nanoparticles-based
nanozymes may help to treat diseases associated with oxidative stress.
78 4 Metal Oxide-Based Nanomaterials for Nanozymes

4.3 Other Metal Oxides

Many other metal oxides have been explored to mimic natural enzymes [35,
136–149].

4.3.1 Vanadium Oxide as Enzyme Mimics

Vanadium oxide nanomaterials (such as V2O5 and V2O3) have been used to mimic
natural enzymes for biosensing, antibiofouling, cytoprotection, etc. [136–139, 150].
Tremel et al. [136] showed that V2O5 nanowires possessed intrinsic peroxidase
mimicking activity. In a subsequent study, they demonstrated that the V2O5
nanowires mimicked natural vanadium haloperoxidase (Fig. 4.18) [137].
A potential mechanism for the vanadium haloperoxidase mimicking activity was
showed in Fig. 4.18b. Interestingly, it has been demonstrated that the nanozymes
could be used to prevent marine biofouling (Fig. 4.18c–f) [137].
Mugesh, D’Silva, and co-workers [138] showed that V2O5 nanowires could also
mimic GSH peroxidase, an important cellular antioxidation enzyme (Fig. 4.19). By
using intracellular glutathione, the nanozyme could protect cells from both intrinsic
and external oxidative injuries. Moreover, the nanozyme has successfully restored
the ROS balance without disturbing the cell’s own antioxidation defense systems.
As shown in Fig. 4.19, a possible mechanism was proposed. H2O2 could be
adsorbed onto the nanozyme and was then reduced into H2O, leading to the for-
mation of complex 1. Complex 1 then interacted with GSH and formed complex 2,
which would be hydrolyzed into complexes 3 and 4. The reaction of complex 4
with H2O2 would regenerate complex 1. It should be noted that complex 3 could be
converted to GSSG, which was further transformed to GSH by glutathione
reductase (GR) and NADPH [138].
DNA and glucose detection has been reported by using polydopamine-coated
V2O5 nanowires as peroxidase mimics [151]. In another report, V2O3-loaded
mesoporous carbon has been used for colorimetric detection of glucose by using its
peroxidase mimicking activity [139].

4.3.2 Cobalt Oxide as Enzyme Mimics

Several reports have showed that cobalt oxide (especially Co3O4) nanomaterials
could mimic peroxidase, catalase, SOD, etc. [150, 152, 153]. For example, Wang
et al. [152] reported the catalase-like activity of Co3O4 nanoparticles. It was pro-
posed that the catalysis could be mediated via a Co2+ → Co3+ → Co2+ regener-
ation mechanism. In a following study, they demonstrated that the Co3O4
nanoparticles’ catalase mimicking activities could be modulated by controlling the
4.3 Other Metal Oxides 79

Fig. 4.18 a TEM image of V2O5 nanowires. b Proposed mechanism for V2O5 nanowires’
vanadium haloperoxidase mimicking activity. c–f Effect of V2O5 nanowires on marine biofouling.
Digital image of a stainless steel plate (2 × 2 cm) covered with a commercially available paint for
boat hulls without and with V2O5 nanowires. The plates were fixed to a boat hull. c, d Immediately
after fixation, both stainless steel plates (with and without V2O5 nanowires) had clean surfaces.
The boat was kept in seawater (lagoon with tidal water directly connected to the Atlantic Ocean).
After 60 days, the boat was taken from the water. e The painted stainless steel plates without V2O5
nanowires suffered from severe natural biofouling. f Plates with V2O5 nanowires showed a
complete absence of biofouling. Adapted from Ref. [137], Copyright 2012, with permission from
Nature Publishing Group
80 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.19 Proposed molecular mechanism for V2O5 nanowires’ GSH peroxidase mimicking
activity. Reprinted from Ref. [138], Copyright 2014, with permission from Nature Publishing
Group

nanoparticles’ morphologies. The nanozymes exhibited different activities in the


order of nanoplates > nanorods > nanocubes. Moreover, they discovered that the
nanozymes’ activities could be specifically enhanced by Ca2+. Based on this phe-
nomenon, they constructed a biosensor for Ca2+. With the biosensor, spiked Ca2+ in
milk was successfully determined [154].
Zhang, Gu, and co-workers [150] found that the Co3O4 nanoparticles possessed
multiple enzyme-like activities (i.e., catalase-like, peroxidase-like, and SOD-like
activities). For all the three mimics, the Co2+ → Co3+ → Co2+ regeneration
mechanism was involved. Compared with Fe3O4 nanoparticles, Co3O4 nanoparti-
cles exhibited higher enzyme mimicking activities. After establishing the enzyme
mimicking activities, they developed an immunoassay for vascular endothelial
growth factor (VEGF) detection [150].
4.3 Other Metal Oxides 81

4.3.3 Copper Oxide as Enzyme Mimics

Chen’s and others’ groups have reported the enzyme mimicking activities of copper
oxide nanomaterials [147, 155–164]. For instance, Chen et al. demonstrated that
CuO nanoparticles with an average size of 30 nm exhibited peroxidase mimicking
activity [155]. Enzyme kinetics study revealed that the CuO-based nanozyme had
higher affinity toward TMB than natural HRP and several other nanozymes (e.g.,
Fe3O4- and FeS-based peroxidase mimics). By combining the CuO nanozyme with
cholesterol oxidase, they have developed a chemiluminescent biosensor for
cholesterol detection [164]. Glucose and lactate detection was also reported by
using CuO nanoparticles as peroxidase mimics [156].

4.3.4 MoO3, TiO2, MnO2, RuO2 as Enzyme Mimics

Tremel et al. [149] studied the enzyme mimicking activities of MoO3 nanoparticles
and found that they could mimic sulfite oxidase, which catalyzed the oxidation of
sulfite to sulfate. After establishing the sulfite oxidase mimicking activity in vitro,
they further demonstrated that the nanozyme could work in living cells and
recovered the sulfite oxidase activity in sulfite oxidase knockdown cells [149]. In
this case, the intracellular sulfite oxidase activity was chemically inhibited by
sodium tungstate treatment. Since native sulfite oxidase is localized within mito-
chondria, the nanozyme was conjugated with triphenylphosphonium moieties for
mitochondria targeting. When the sulfite oxidase deficient cells were treated with
the nanozymes, their sulfite oxidase activity was restored [149].
Dong et al. fabricated TiO2 nanotube arrays via a potentiostatic anodization
strategy. They further showed that the TiO2 nanotube arrays exhibited
peroxidase-like activity. Sensitive detection of H2O2 was demonstrated by using the
TiO2 nanotube arrays as a working electrode, which showed catalytic activity toward
H2O2 reduction. Moreover, when GOx was immobilized onto the TiO2 nanotube
arrays, an electrochemical sensor for glucose detection was obtained. As shown in
Fig. 4.20, the biosensor exhibited excellent sensitivity and selectivity toward glu-
cose determination. Besides, the biosensor was further used to measure the glucose
concentrations in diabetes patients, showing satisfactory performance [148].
Gao et al. [140] have constructed an electrochemical DNA biosensor by using
the peroxidase mimicking activity of RuO2 nanoparticles. When the RuO2
nanoparticles labeled probe ssDNA was specifically assembled onto an electrode in
a sandwich manner with the capture ssDNA and target ssDNA, the nanoparticles
catalyzed the deposition of polyaniline and produced catalytic electrochemical
signals for detection. Zhang and co-workers demonstrated that MnO2 nanowires
were excellent peroxidase mimics. They then developed an immunoassay to detect
sulfate-reducing bacteria [144].
82 4 Metal Oxide-Based Nanomaterials for Nanozymes

Fig. 4.20 a Cyclic voltammograms (CVs) obtained at the GOx/TiO2 nanotube array electrode in
0.1 M pH 5.5 PBS without (a) and with (b) 4 mM glucose. b CVs obtained at the TiO2 nanotube
array electrode in 0.1 M pH 5.5 PBS without (a) and with (b) 4 mM glucose. c Amperometric
response obtained at the GOD/TiO2 nanotube array electrode in 0.1 M pH 5.5 PBS upon
successive injection of 0.2 mM glucose for each step at −0.35 V, the inset is the linear fitted
current plot to glucose concentration. d Amperometric response at the GOD/TiO2 nanotube array
electrode in 0.1 M pH 5.5 PBS at −0.35 V with sequential injection of 1 mM glucose (a), 2 mM
fructose (b), 2 mM lactose (c), 2 mM sucrose (d), 2 mM maltose (e), and 1 mM glucose (f).
Reprinted from Ref. [148], Copyright 2013, with permission from Royal Society of Chemistry

References

1. Esch, F., Fabris, S., Zhou, L., Montini, T., Africh, C., Fornasiero, P., et al. (2005). Electron
localization determines defect formation on ceria substrates. Science, 309, 752–755.
2. Karakoti, A., Singh, S., Dowding, J. M., Seal, S., & Self, W. T. (2010). Redox-active radical
scavenging nanomaterials. Chemical Society Reviews, 39, 4422–4432.
3. Grulke, E., Reed, K., Beck, M., Huang, X., Cormack, A., & Seal, S. (2014). Nanoceria:
Factors affecting its pro- and anti-oxidant properties. Environmental Science-Nano, 1,
429–444.
4. Kumar, A., Das, S., Munusamy, P., Self, W., Baer, D. R., Sayle, D. C., et al. (2014).
Behavior of nanoceria in biologically-relevant environments. Environmental Science-Nano,
1, 516–532.
5. Zhao, H., Dong, Y. M., Jiang, P. P., Wang, G. L., & Zhang, J. J. (2015). Highly dispersed
CeO2 on TiO2 nanotube: A synergistic nanocomposite with superior peroxidase-like activity.
ACS Applied Materials & Interfaces, 7, 6451–6461.
References 83

6. Wei, H., & Wang, E. K. (2013). Nanomaterials with enzyme-like characteristics


(nanozymes): Next-generation artificial enzymes. Chemical Society Reviews, 42, 6060–6093.
7. Wang, X. Y., Hu, Y. H., & Wei, H. (2016). Nanozymes in bionanotechnology: From sensing
to therapeutics and beyond. Inorganic Chemistry Frontiers, 3, 41–60.
8. Xu, C., & Qu, X. G. (2014). Cerium oxide nanoparticle: A remarkably versatile rare earth
nanomaterial for biological applications. NPG Asia Materials, 6, e60.
9. Tarnuzzer, R. W., Colon, J., Patil, S., & Seal, S. (2005). Vacancy engineered ceria
nanostructures for protection from radiation-induced cellular damage. Nano Letters, 5,
2573–2577.
10. Karakoti, A. S., Singh, S., Kumar, A., Malinska, M., Kuchibhatla, S., Wozniak, K., et al.
(2009). PEGylated nanoceria as radical scavenger with tunable redox chemistry. Journal of
the American Chemical Society, 131, 14144–14145.
11. Liu, X. Y., Wei, W., Yuan, Q., Zhang, X., Li, N., Du, Y. G., et al. (2012). Apoferritin-CeO2
nano-truffle that has excellent artificial redox enzyme activity. Chemical Communications,
48, 3155–3157.
12. Singh, S., Dosani, T., Karakoti, A. S., Kumar, A., Seal, S., & Self, W. T. (2011).
A phosphate-dependent shift in redox state of cerium oxide nanoparticles and its effects on
catalytic properties. Biomaterials, 32, 6745–6753.
13. Kamada, K., & Soh, N. (2015). Enzyme-mimetic activity of Ce-intercalated titanate
nanosheets. Journal of Physical Chemistry B, 119, 5309–5314.
14. McCormack, R. N., Mendez, P., Barkam, S., Neal, C. J., Das, S., & Seal, S. (2014).
Inhibition of nanoceria’s catalytic activity due to Ce3+ site-specific interaction with phosphate
ions. Journal of Physical Chemistry C, 118, 18992–19006.
15. Cafun, J.-D., Kvashnina, K. O., Casals, E., Puntes, V. F., & Glatzel, P. (2013). Absence of
Ce3+ sites in chemically active colloidal ceria nanoparticles. ACS Nano, 7, 10726–10732.
16. Babu, S., Cho, J. H., Dowding, J. M., Heckert, E., Komanski, C., Das, S., et al. (2010).
Multicolored redox active upconverter cerium oxide nanoparticle for bio-imaging and
therapeutics. Chemical Communications, 46, 6915–6917.
17. Madero-Visbal, R. A., Alvarado, B. E., Colon, J. F., Baker, C. H., Wason, M. S., Isley, B.,
et al. (2012). Harnessing nanoparticles to improve toxicity after head and neck radiation.
Nanomedicine-Nanotechnology Biology and Medicine, 8, 1223–1231.
18. Kuchma, M. H., Komanski, C. B., Colon, J., Teblum, A., Masunov, A. E., Alvarado, B.,
et al. (2010). Phosphate ester hydrolysis of biologically relevant molecules by cerium oxide
nanoparticles. Nanomedicine-Nanotechnology Biology and Medicine, 6, 738–744.
19. Colon, J., Hsieh, N., Ferguson, A., Kupelian, P., Seal, S., Jenkins, D. W., et al. (2010).
Cerium oxide nanoparticles protect gastrointestinal epithelium from radiation-induced
damage by reduction of reactive oxygen species and upregulation of superoxide dismutase 2.
Nanomedicine-Nanotechnology Biology and Medicine, 6, 698–705.
20. Hayat, A., Bulbul, G., & Andreescu, S. (2014). Probing phosphatase activity using redox
active nanoparticles: A novel colorimetric approach for the detection of enzyme activity.
Biosensors & Bioelectronics, 56, 334–339.
21. Oezel, R. E., Alkasir, R. S. J., Ray, K., Wallace, K. N., & Andreescu, S. (2013). Comparative
evaluation of intestinal nitric oxide in embryonic zebrafish exposed to metal oxide
nanoparticles. Small (Weinheim an der Bergstrasse, Germany), 9, 4250–4261.
22. Tsai, Y.-Y., Oca-Cossio, J., Lin, S.-M., Woan, K., Yu, P.-C., & Sigmund, W. (2008).
Reactive oxygen species scavenging properties of ZrO2-CeO2 solid solution nanoparticles.
Nanomedicine, 3, 637–645.
23. Das, S., Dowding, J. M., Klump, K. E., McGinnis, J. F., Self, W., & Seal, S. (2013). Cerium
oxide nanoparticles: applications and prospects in nanomedicine. Nanomedicine, 8,
1483–1508.
24. Karakoti, A. S., Monteiro-Riviere, N. A., Aggarwal, R., Davis, J. P., Narayan, R. J., Self, W.
T., et al. (2008). Nanoceria as antioxidant: Synthesis and biomedical applications. Journal of
the Minerals Metals and Materials Society, 60, 33–37.
84 4 Metal Oxide-Based Nanomaterials for Nanozymes

25. Bhushan, B., & Gopinath, P. (2015). Antioxidant nanozyme: A facile synthesis and
evaluation of the reactive oxygen species scavenging potential of nanoceria encapsulated
albumin nanoparticles. Journal of Materials Chemistry B, 3, 4843–4852.
26. Li, M., Shi, P., Xu, C., Ren, J., & Qu, X. (2013). Cerium oxide caged metal chelator:
Anti-aggregation and anti-oxidation integrated H2O2-responsive controlled drug release for
potential Alzheimer’s disease treatment. Chemical Science, 4, 2536–2542.
27. Li, H., Yang, Z.-Y., Liu, C., Zeng, Y.-P., Hao, Y.-H., Gu, Y., et al. (2015). PEGylated ceria
nanoparticles used for radioprotection on human liver cells under gamma-ray irradiation.
Free Radical Biology and Medicine, 87, 26–35.
28. Muhammad, F., Wang, A., Qi, W., Zhang, S., & Zhu, G. (2014). Intracellular antioxidants
dissolve man-made antioxidant nanoparticles: Using redox vulnerability of nanoceria to
develop a responsive drug delivery system. ACS Applied Materials & Interfaces, 6,
19424–19433.
29. Weaver, J. D., & Stabler, C. L. (2015). Antioxidant cerium oxide nanoparticle hydrogels for
cellular encapsulation. Acta Biomaterialia, 16, 136–144.
30. Xu, C., Liu, Z., Wu, L., Ren, J., & Qu, X. (2014). Nucleoside triphosphates as promoters to
enhance nanoceria enzyme-like activity and for single-nucleotide polymorphism typing.
Advanced Functional Materials, 24, 1624–1630.
31. Xu, C., Lin, Y., Wang, J., Wu, L., Wei, W., Ren, J., et al. (2013). Nanoceria-triggered
synergetic drug release based on CeO2-capped mesoporous silica host-guest interactions and
switchable enzymatic activity and cellular effects of CeO2. Advanced Healthcare Materials,
2, 1591–1599.
32. Patil, S., Sandberg, A., Heckert, E., Self, W., & Seal, S. (2007). Protein adsorption and
cellular uptake of cerium oxide nanoparticles as a function of zeta potential. Biomaterials, 28,
4600–4607.
33. Xue, Y., Luan, Q., Yang, D., Yao, X., & Zhou, K. (2011). Direct evidence for hydroxyl
radical scavenging activity of cerium oxide nanoparticles. Journal of Physical Chemistry C,
115, 4433–4438.
34. Barkam, S., Das, S., Saraf, S., McCormack, R., Richardson, D., Atencio, L., et al. (2015).
The change in antioxidant properties of dextran-coated redox active nanoparticles due to
synergetic photoreduction-oxidation. Chemistry-A European Journal, 21, 12646–12656.
35. Schubert, D., Dargusch, R., Raitano, J., & Chan, S. W. (2006). Cerium and yttrium oxide
nanoparticles are neuroprotective. Biochemical and Biophysical Research Communications,
342, 86–91.
36. Korsvik, C., Patil, S., Seal, S., & Self, W. T. (2007). Superoxide dismutase mimetic
properties exhibited by vacancy engineered ceria nanoparticles. Chemical Communications,
1056–1058.
37. Heckert, E. G., Karakoti, A. S., Seal, S., & Self, W. T. (2008). The role of cerium redox state
in the SOD mimetic activity of nanoceria. Biomaterials, 29, 2705–2709.
38. Sardesai, N. P., Andreescu, D., & Andreescu, S. (2013). Electroanalytical evaluation of
antioxidant activity of cerium oxide nanoparticles by nanoparticle collisions at
microelectrodes. Journal of the American Chemical Society, 135, 16770–16773.
39. Celardo, I., Pedersen, J. Z., Traversa, E., & Ghibelli, L. (2011). Pharmacological potential of
cerium oxide nanoparticles. Nanoscale, 3, 1411–1420.
40. Li, Y., He, X., Yin, J.-J., Ma, Y., Zhang, P., Li, J., et al. (2015). Acquired
superoxide-scavenging ability of ceria nanoparticles. Angewandte Chemie-International
Edition, 54, 1832–1835.
41. Celardo, I., De Nicola, M., Mandoli, C., Pedersen, J. Z., Traversa, E., & Ghibelli, L. (2011).
Ce3+ ions determine redox-dependent anti-apoptotic effect of cerium oxide nanoparticles.
ACS Nano, 5, 4537–4549.
42. Fernandez-Garcia, S., Jiang, L., Tinoco, M., Hungria, A. B., Han, J., Blanco, G., et al.
(2016). Enhanced hydroxyl radical scavenging activity by doping lanthanum in ceria
nanocubes. Journal of Physical Chemistry C, 120, 1891–1901.
References 85

43. Zhu, A. P., Sun, K., & Petty, H. R. (2012). Titanium doping reduces superoxide dismutase
activity, but not oxidase activity, of catalytic CeO2 nanoparticles. Inorganic Chemistry
Communications, 15, 235–237.
44. Lin, Y., Xu, C., Ren, J., & Qu, X. (2012). Using thermally regenerable cerium oxide
nanoparticles in biocomputing to perform label-free, resettable, and colorimetric logic
operations. Angewandte Chemie-International Edition, 51, 12579–12583.
45. Lord, M. S., Jung, M., Teoh, W. Y., Gunawan, C., Vassie, J. A., Amal, R., et al. (2012).
Cellular uptake and reactive oxygen species modulation of cerium oxide nanoparticles in
human monocyte cell line U937. Biomaterials, 33, 7915–7924.
46. Selvaraj, V., Nepal, N., Rogers, S., Manne, N. D. P. K., Arvapalli, R., Rice, K. M., et al.
(2015). Inhibition of MAP kinase/NF-kB mediated signaling and attenuation of
lipopolysaccharide induced severe sepsis by cerium oxide nanoparticles. Biomaterials, 59,
160–171.
47. Chen, J. P., Patil, S., Seal, S., & McGinnis, J. F. (2006). Rare earth nanoparticles prevent
retinal degeneration induced by intracellular peroxides. Nature Nanotechnology, 1, 142–150.
48. Cai, X., Sezate, S. A., Seal, S., & McGinnis, J. F. (2012). Sustained protection against
photoreceptor degeneration in tubby mice by intravitreal injection of nanoceria. Biomaterials,
33, 8771–8781.
49. Cai, X., Seal, S., & McGinnis, J. F. (2014). Sustained inhibition of neovascularization in
vldlr–/– mice following intravitreal injection of cerium oxide nanoparticles and the role of the
ASK1-P38/JNK-NF-kappa B pathway. Biomaterials, 35, 249–258.
50. Wong, L. L., Pye, Q. N., Chen, L., Seal, S., & McGinnis, J. F. (2015). Defining the catalytic
activity of nanoceria in the p23h-1 rat, a photoreceptor degeneration model. PLoS ONE, 10,
e0121977.
51. Das, M., Patil, S., Bhargava, N., Kang, J. F., Riedel, L. M., Seal, S., et al. (2007).
Auto-catalytic ceria nanoparticles offer neuroprotection to adult rat spinal cord neurons.
Biomaterials, 28, 1918–1925.
52. Estevez, A. Y., Pritchard, S., Harper, K., Aston, J. W., Lynch, A., Lucky, J. J., et al. (2011).
Neuroprotective mechanisms of cerium oxide nanoparticles in a mouse hippocampal brain
slice model of ischemia. Free Radical Biology and Medicine, 51, 1155–1163.
53. Kim, C. K., Kim, T., Choi, I.-Y., Soh, M., Kim, D., Kim, Y.-J., et al. (2012). Ceria
nanoparticles that can protect against ischemic stroke. Angewandte Chemie-International
Edition, 51, 11039–11043.
54. Heckman, K. L., DeCoteau, W., Estevez, A., Reed, K. J., Costanzo, W., Sanford, D., et al.
(2013). Custom cerium oxide nanoparticles protect against a free radical mediated
autoimmune degenerative disease in the brain. ACS Nano, 7, 10582–10596.
55. Kwon, H. J., Cha, M.-Y., Kim, D., Kim, D. K., Soh, M., Shin, K., et al. (2016).
Mitochondria-targeting ceria nanoparticles as antioxidants for alzheimer’s disease. ACS
Nano, 10, 2860–2870.
56. Pagliari, F., Mandoli, C., Forte, G., Magnani, E., Pagliari, S., Nardone, G., et al. (2012).
Cerium oxide nanoparticles protect cardiac progenitor cells from oxidative stress. ACS Nano,
6, 3767–3775.
57. Niu, J., Azfer, A., Rogers, L. M., Wang, X., & Kolattukudy, P. E. (2007). Cardioprotective
effects of cerium oxide nanoparticles in a transgenic murine model of cardiomyopathy.
Cardiovascular Research, 73, 549–559.
58. Kolli, M. B., Manne, N. D. P. K., Para, R., Nalabotu, S. K., Nandyala, G., Shokuhfar, T.,
et al. (2014). Cerium oxide nanoparticles attenuate monocrotaline induced right ventricular
hypertrophy following pulmonary arterial hypertension. Biomaterials, 35, 9951–9962.
59. Niu, J. L., Azfer, A., Rogers, L. M., Wang, X. H., & Kolattukudy, P. E. (2007).
Cardioprotective effects of cerium oxide nanoparticles in a transgenic murine model of
cardiomyopathy. Cardiovascular Research, 73, 549–559.
60. Alili, L., Sack, M., Karakoti, A. S., Teuber, S., Puschmann, K., Hirst, S. M., et al. (2011).
Combined cytotoxic and anti-invasive properties of redox-active nanoparticles in
tumor-stroma interactions. Biomaterials, 32, 2918–2929.
86 4 Metal Oxide-Based Nanomaterials for Nanozymes

61. Alili, L., Sack, M., von Montfort, C., Giri, S., Das, S., Carroll, K. S., et al. (2013).
Downregulation of tumor growth and invasion by redox-active nanoparticles. Antioxidants &
Redox Signaling, 19, 765–778.
62. Giri, S., Karakoti, A., Graham, R. P., Maguire, J. L., Reilly, C. M., Seal, S., et al. (2013).
Nanoceria: A rare-earth nanoparticle as a novel anti-angiogenic therapeutic agent in ovarian
cancer. PLoS ONE, 8, e54578.
63. Alpaslan, E., Yazici, H., Golshan, N. H., Ziemer, K. S., & Webster, T. J. (2015).
pH-dependent activity of dextran-coated cerium oxide nanoparticles on prohibiting
osteosarcoma cell proliferation. ACS Biomaterials-Science & Engineering, 1, 1096–1103.
64. Mandoli, C., Pagliari, F., Pagliari, S., Forte, G., Di Nardo, P., Licoccia, S., et al. (2010). Stem
cell aligned growth induced by CeO2 nanoparticles in PLGA scaffolds with improved
bioactivity for regenerative medicine. Advanced Functional Materials, 20, 1617–1624.
65. Naganuma, T., & Traversa, E. (2014). The effect of cerium valence states at cerium oxide
nanoparticle surfaces on cell proliferation. Biomaterials, 35, 4441–4453.
66. Chigurupati, S., Mughal, M. R., Okun, E., Das, S., Kumar, A., McCaffery, M., et al. (2013).
Effects of cerium oxide nanoparticles on the growth of keratinocytes, fibroblasts and vascular
endothelial cells in cutaneous wound healing. Biomaterials, 34, 2194–2201.
67. Karakoti, A. S., Tsigkou, O., Yue, S., Lee, P. D., Stevens, M. M., Jones, J. R., et al. (2010).
Rare earth oxides as nanoadditives in 3-D nanocomposite scaffolds for bone regeneration.
Journal of Materials Chemistry, 20, 8912–8919.
68. Ponnurangam, S., O’Connell, G. D., Chernyshova, I. V., Wood, K., Hung, C. T.-H., &
Somasundaran, P. (2014). Beneficial effects of cerium oxide nanoparticles in development of
chondrocyte-seeded hydrogel constructs and cellular response to interleukin insults. Tissue
Engineering Part A, 20, 2908–2919.
69. Pulido-Reyes, G., Rodea-Palomares, I., Das, S., Sakthivel, T. S., Leganes, F., Rosal, R., et al.
(2015). Untangling the biological effects of cerium oxide nanoparticles: the role of surface
valence states. Scientific Reports, 5, 15613.
70. Hirst, S. M., Karakoti, A. S., Tyler, R. D., Sriranganathan, N., Seal, S., & Reilly, C. M.
(2009). Anti-inflammatory properties of cerium oxide nanoparticles. Small (Weinheim an der
Bergstrasse, Germany), 5, 2848–2856.
71. Oro, D., Yudina, T., Fernandez-Varo, G., Casals, E., Reichenbach, V., Casals, G., et al.
(2016). Cerium oxide nanoparticles reduce steatosis, portal hypertension and display
anti-inflammatory properties in rats with liver fibrosis. Journal of Hepatology, 64, 691–698.
72. Perez, J. M., Asati, A., Nath, S., & Kaittanis, C. (2008). Synthesis of biocompatible
dextran-coated nanoceria with pH-dependent antioxidant properties. Small (Weinheim an der
Bergstrasse, Germany), 4, 552–556.
73. Lee, S. S., Song, W., Cho, M., Puppala, H. L., Phuc, N., Zhu, H., et al. (2013). Antioxidant
properties of cerium oxide nanocrystals as a function of nanocrystal diameter and surface
coating. ACS Nano, 7, 9693–9703.
74. Pelletier, D. A., Suresh, A. K., Holton, G. A., McKeown, C. K., Wang, W., Gu, B., et al.
(2010). Effects of engineered cerium oxide nanoparticles on bacterial growth and viability.
Applied and Environmental Microbiology, 76, 7981–7989.
75. Shah, V., Shah, S., Shah, H., Rispoli, F. J., McDonnell, K. T., Workeneh, S., et al. (2012).
Antibacterial activity of polymer coated cerium oxide nanoparticles. PLoS ONE, 7,
e0047827.
76. Son, D., Lee, J., Lee, D. J., Ghaffari, R., Yun, S., Kim, S. J., et al. (2015). Bioresorbable
electronic stent integrated with therapeutic nanoparticles for endovascular diseases. ACS
Nano, 9, 5937–5946.
77. Liu, B. W., Sun, Z. Y., Huang, P. J. J., & Liu, J. W. (2015). Hydrogen peroxide displacing
DNA from nanoceria: Mechanism and detection of glucose in serum. Journal of the
American Chemical Society, 137, 1290–1295.
78. Pirmohamed, T., Dowding, J. M., Singh, S., Wasserman, B., Heckert, E., Karakoti, A. S.,
et al. (2010). Nanoceria exhibit redox state-dependent catalase mimetic activity. Chemical
Communications, 46, 2736–2738.
References 87

79. Nicolini, V., Gambuzzi, E., Malavasi, G., Menabue, L., Menziani, M. C., Lusvardi, G., et al.
(2015). Evidence of catalase mimetic activity in Ce3+/Ce4+ doped bioactive glasses. Journal
of Physical Chemistry B, 119, 4009–4019.
80. Gao, W., Wei, X., Wang, X., Cui, G., Liu, Z., & Tang, B. (2016). A competitive
coordination-based CeO2 nanowire-DNA nanosensor: fast and selective detection of
hydrogen peroxide in living cells and in vivo. Chemical Communications, 52, 3643–3646.
81. Tian, Z. M., Li, J., Zhang, Z. Y., Gao, W., Zhou, X. M., & Qu, Y. Q. (2015). Highly sensitive
and robust peroxidase-like activity of porous nanorods of ceria and their application for
breast cancer detection. Biomaterials, 59, 116–124.
82. Artiglia, L., Agnoli, S., Paganini, M. C., Cattelan, M., & Granozzi, G. (2014). TiO2@CeOx
core-shell nanoparticles as artificial enzymes with peroxidase-like activity. ACS Applied
Materials & Interfaces, 6, 20130–20136.
83. Asati, A., Santra, S., Kaittanis, C., Nath, S., & Perez, J. M. (2009). Oxidase-like activity of
polymer-coated cerium oxide nanoparticles. Angewandte Chemie-International Edition, 48,
2308–2312.
84. Pautler, R., Kelly, E. Y., Huang, P. J. J., Cao, J., Liu, B. W., & Liu, J. W. (2013).
Attaching DNA to nanoceria: Regulating oxidase activity and fluorescence quenching. ACS
Applied Materials & Interfaces, 5, 6820–6825.
85. Kim, M. I., Park, K. S., & Park, H. G. (2014). Ultrafast colorimetric detection of nucleic
acids based on the inhibition of the oxidase activity of cerium oxide nanoparticles. Chemical
Communications, 50, 9577–9580.
86. Dowding, J. M., Dosani, T., Kumar, A., Seal, S., & Self, W. T. (2012). Cerium oxide
nanoparticles scavenge nitric oxide radical (∙NO). Chemical Communications, 48,
4896–4898.
87. Vernekar, A. A., Das, T., & Mugesh, G. (2016). Vacancy-engineered nanoceria: Enzyme
mimetic hotspots for the degradation of nerve agents. Angewandte Chemie-International
Edition, 55, 1412–1416.
88. Gao, L. Z., Zhuang, J., Nie, L., Zhang, J. B., Zhang, Y., Gu, N., et al. (2007). Intrinsic
peroxidase-like activity of ferromagnetic nanoparticles. Nature Nanotechnology, 2, 577–583.
89. Wang, L. J., Min, Y., Xu, D. D., Yu, F. J., Zhou, W. Z., & Cuschieri, A. (2014). Membrane
lipid peroxidation by the peroxidase-like activity of magnetite nanoparticles. Chemical
Communications, 50, 11147–11150.
90. Kim, M. I., Shim, J., Li, T., Lee, J., & Park, H. G. (2011). Fabrication of nanoporous
nanocomposites entrapping Fe3O4 magnetic nanoparticles and oxidases for colorimetric
biosensing. Chemistry-A European Journal, 17, 10700–10707.
91. Liu, C. H., & Tseng, W. L. (2011). Oxidase-functionalized Fe3O4 nanoparticles for
fluorescence sensing of specific substrate. Analytica Chimica Acta, 703, 87–93.
92. Kim, M. I., Ye, Y., Won, B. Y., Shin, S., Lee, J., & Park, H. G. (2011). A highly efficient
electrochemical biosensing platform by employing conductive nanocomposite entrapping
magnetic nanoparticles and oxidase in mesoporous carbon foam. Advanced Functional
Materials, 21, 2868–2875.
93. Liu, S. H., Lu, F., Xing, R. M., & Zhu, J. J. (2011). Structural effects of Fe3O4 nanocrystals
on peroxidase-like activity. Chemistry-A European Journal, 17, 620–625.
94. Wu, Y. H., Song, M. J., Xin, Z. A., Zhang, X. Q., Zhang, Y., Wang, C. Y., et al. (2011).
Ultra-small particles of iron oxide as peroxidase for immunohistochemical detection.
Nanotechnology, 22, 225703.
95. Dong, Y. L., Zhang, H. G., Rahman, Z. U., Su, L., Chen, X. J., Hu, J., et al. (2012). Graphene
oxide-Fe3O4 magnetic nanocomposites with peroxidase-like activity for colorimetric
detection of glucose. Nanoscale, 4, 3969–3976.
96. Kim, M. I., Shim, J., Li, T., Woo, M. A., Cho, D., Lee, J., et al. (2012). Colorimetric
quantification of galactose using a nanostructured multi-catalyst system entrapping galactose
oxidase and magnetic nanoparticles as peroxidase mimetics. Analyst, 137, 1137–1143.
88 4 Metal Oxide-Based Nanomaterials for Nanozymes

97. Lee, J. W., Jeon, H. J., Shin, H. J., & Kang, J. K. (2012). Superparamagnetic Fe3O4
nanoparticles-carbon nitride nanotube hybrids for highly efficient peroxidase mimetic
catalysts. Chemical Communications, 48, 422–424.
98. Ma, Y. H., Zhang, Z. Y., Ren, C. L., Liu, G. Y., & Chen, X. G. (2012). A novel colorimetric
determination of reduced glutathione in A549 cells based on Fe3O4 magnetic nanoparticles as
peroxidase mimetics. Analyst, 137, 485–489.
99. Fan, K. L., Cao, C. Q., Pan, Y. X., Lu, D., Yang, D. L., Feng, J., et al. (2012).
Magnetoferritin nanoparticles for targeting and visualizing tumour tissues. Nature
Nanotechnology, 7, 459–464.
100. Zhang, X. Q., Gong, S. W., Zhang, Y., Yang, T., Wang, C. Y., & Gu, N. (2010). Prussian
blue modified iron oxide magnetic nanoparticles and their high peroxidase-like activity.
Journal of Materials Chemistry, 20, 5110–5116.
101. Chaudhari, K. N., Chaudhari, N. K., & Yu, J.-S. (2012). Peroxidase mimic activity of
hematite iron oxides (α-Fe2O3) with different nanostructres. Catalysis Science & Technology,
2, 119–124.
102. Guo, F. F., Yang, W., Jiang, W., Geng, S., Peng, T., & Li, J. L. (2012). Magnetosomes
eliminate intracellular reactive oxygen species in Magnetospirillum gryphiswaldense MSR-1.
Environmental Microbiology, 14, 1722–1729.
103. Fan, Y. W., & Huang, Y. M. (2012). The effective peroxidase-like activity of
chitosan-functionalized CoFe2O4 nanoparticles for chemiluminescence sensing of
hydrogen peroxide and glucose. Analyst, 137, 1225–1231.
104. Shi, W. B., Wang, H., & Huang, Y. M. (2011). Luminol-silver nitrate chemiluminescence
enhancement induced by cobalt ferrite nanoparticles. Luminescence, 26, 547–552.
105. Zhang, X. D., He, S. H., Chen, Z. H., & Huang, Y. M. (2013). CoFe2O4 nanoparticles as
oxidase mimic-mediated chemiluminescence of aqueous luminol for sulfite in white wines.
Journal of Agricultural and Food Chemistry, 61, 840–847.
106. Luo, W., Li, Y. S., Yuan, J., Zhu, L. H., Liu, Z. D., Tang, H. Q., et al. (2010). Ultrasensitive
fluorometric determination of hydrogen peroxide and glucose by using multiferroic BiFeO3
nanoparticles as a catalyst. Talanta, 81, 901–907.
107. Bhattacharya, D., Baksi, A., Banerjee, I., Ananthakrishnan, R., Maiti, T. K., & Pramanik,
P. (2011). Development of phosphonate modified Fe(1−x)MnxFe2O4 mixed ferrite
nanoparticles: Novel peroxidase mimetics in enzyme linked immunosorbent assay.
Talanta, 86, 337–348.
108. Yan, L., Liu, X., Liu, W.-X., Tan, X.-Q., Xiong, F., Gu, N., et al. (2015). Fe2O3
nanoparticles suppress Kv1.3 channels via affecting the redox activity of Kv2 subunit in
Jurkat T cells. Nanotechnology, 26, 505103.
109. Lu, C., Liu, X., Li, Y., Yu, F., Tang, L., Hu, Y., et al. (2015). Multifunctional janus hematite
silica nanoparticles: Mimicking peroxidase-like activity and sensitive colorimetric detection
of glucose. ACS Applied Materials & Interfaces, 7, 15395–15402.
110. Kim, J. A., Lee, N. H., Kim, B. H., Rhee, W. J., Yoon, S., Hyeon, T., et al. (2011).
Enhancement of neurite outgrowth in PC12 cells by iron oxide nanoparticles. Biomaterials,
32, 2871–2877.
111. He, X. L., Tan, L. F., Chen, D., Wu, X. L., Ren, X. L., Zhang, Y. Q., et al. (2013). Fe3O4–
Au@mesoporous SiO2 microspheres: An ideal artificial enzymatic cascade system. Chemical
Communications, 49, 4643–4645.
112. An, Q., Sun, C. Y., Li, D., Xu, K., Guo, J., & Wang, C. C. (2013). Peroxidase-like activity of
Fe3O4@carbon nanoparticles enhances ascorbic acid-induced oxidative stress and selective
damage to PC-3 prostate cancer cells. ACS Applied Materials & Interfaces, 5, 13248–13257.
113. Gao, L. Z., Giglio, K. M., Nelson, J. L., Sondermann, H., & Travis, A. J. (2014).
Ferromagnetic nanoparticles with peroxidase-like activity enhance the cleavage of biological
macromolecules for biofilm elimination. Nanoscale, 6, 2588–2593.
114. Su, L., Feng, J., Zhou, X. M., Ren, C. L., Li, H. H., & Chen, X. G. (2012). Colorimetric
detection of urine glucose based ZnFe2O4 magnetic nanoparticles. Analytical Chemistry, 84,
5753–5758.
References 89

115. Li, Q., Tang, G. G., Xiong, X. W., Cao, Y. L., Chen, L. L., Xu, F. G., et al. (2015). Carbon
coated magnetite nanoparticles with improved water-dispersion and peroxidase-like activity
for colorimetric sensing of glucose. Sensors and Actuators B-Chemical, 215, 86–92.
116. Shin, S., Yoon, H., & Jang, J. (2008). Polymer-encapsulated iron oxide nanoparticles as
highly efficient Fenton catalysts. Catalysis Communications, 10, 178–182.
117. Wang, N., Zhu, L. H., Wang, D. L., Wang, M. Q., Lin, Z. F., & Tang, H. Q. (2010).
Sono-assisted preparation of highly-efficient peroxidase-like Fe3O4 magnetic nanoparticles
for catalytic removal of organic pollutants with H2O2. Ultrasonics Sonochemistry, 17,
526–533.
118. Niu, H. Y., Zhang, D., Zhang, S. X., Zhang, X. L., Meng, Z. F., & Cai, Y. Q. (2011). Humic
acid coated Fe3O4 magnetic nanoparticles as highly efficient Fenton-like catalyst for
complete mineralization of sulfathiazole. Journal of Hazardous Materials, 190, 559–565.
119. Wei, H., & Wang, E. (2008). Fe3O4 magnetic nanoparticles as peroxidase mimetics and their
applications in H2O2 and glucose detection. Analytical Chemistry, 80, 2250–2254.
120. Wang, X. Y., Hu, Y. H., & Wei, H. (2016). Nanozymes in bionanotechnology: From sensing
to therapeutics and beyond. Inorganic Chemistry Frontiers, 3, 41–60.
121. Liang, M. M., Fan, K. L., Pan, Y., Jiang, H., Wang, F., Yang, D. L., et al. (2013). Fe3O4
magnetic nanoparticle peroxidase mimetic-based colorimetric assay for the rapid detection of
organophosphorus pesticide and nerve agent. Analytical Chemistry, 85, 308–312.
122. Park, K. S., Kim, M. I., Cho, D. Y., & Park, H. G. (2011). Label-free colorimetric detection
of nucleic acids based on target-induced shielding against the peroxidase-mimicking activity
of magnetic nanoparticles. Small (Weinheim an der Bergstrasse, Germany), 7, 1521–1525.
123. Thiramanas, R., Jangpatarapongsa, K., Tangboriboonrat, P., & Polpanich, D. (2013).
Detection of vibrio cholerae using the intrinsic catalytic activity of a magnetic polymeric
nanoparticle. Analytical Chemistry, 85, 5996–6002.
124. Liu, B. W., & Liu, J. W. (2015). Accelerating peroxidase mimicking nanozymes using DNA.
Nanoscale, 7, 13831–13835.
125. Duan, D., Fan, K., Zhang, D., Tan, S., Liang, M., Liu, Y., et al. (2015). Nanozyme-strip for
rapid local diagnosis of Ebola. Biosensors & Bioelectronics, 74, 134–141.
126. Gao, L. Z., Wu, J. M., Lyle, S., Zehr, K., Cao, L. L., & Gao, D. (2008). Magnetite
nanoparticle-linked immunosorbent assay. Journal of Physical Chemistry C, 112,
17357–17361.
127. Kaittanis, C., Santra, S., & Perez, J. M. (2009). Role of nanoparticle valency in the
nondestructive magnetic-relaxation-mediated detection and magnetic isolation of cells in
complex media. Journal of the American Chemical Society, 131, 12780–12791.
128. Zhang, Z. X., Wang, Z. J., Wang, X. L., & Yang, X. R. (2010). Magnetic nanoparticle-linked
colorimetric aptasensor for the detection of thrombin. Sensors and Actuators B-Chemical,
147, 428–433.
129. Zheng, T. T., Zhang, Q. F., Feng, S., Zhu, J. J., Wang, Q., & Wang, H. (2014). Robust
nonenzymatic hybrid nanoelectrocatalysts for signal amplification toward ultrasensitive
electrochemical cytosensing. Journal of the American Chemical Society, 136, 2288–2291.
130. Zhang, S., Zhou, G. L., Xu, X. L., Cao, L. L., Liang, G. H., Chen, H., et al. (2011).
Development of an electrochemical aptamer-based sensor with a sensitive Fe3O4
nanoparticle-redox tag for reagentless protein detection. Electrochemistry Communications,
13, 928–931.
131. Chen, Z. W., Yin, J. J., Zhou, Y. T., Zhang, Y., Song, L., Song, M. J., et al. (2012). Dual
enzyme-like activities of iron oxide nanoparticles and their implication for diminishing
cytotoxicity. ACS Nano, 6, 4001–4012.
132. Zhang, Y., Wang, Z. Y., Li, X. J., Wang, L., Yin, M., Wang, L. H., et al. (2016). Dietary iron
oxide nanoparticles delay aging and ameliorate neurodegeneration in drosophila. Advanced
Materials, 28, 1387–1393.
133. Xiong, F., Wang, H., Feng, Y. D., Li, Y. M., Hua, X. Q., Pang, X. Y., et al. (2015).
Cardioprotective activity of iron oxide nanoparticles. Scientific Reports, 5, 8579.
90 4 Metal Oxide-Based Nanomaterials for Nanozymes

134. Li, Q., Tang, G., Xue, S., He, X., Miao, P., Li, Y., et al. (2013). Silica-coated
superparamagnetic iron oxide nanoparticles targeting of EPCs in ischemic brain injury.
Biomaterials, 34, 4982–4992.
135. Wang, Q., Chen, B., Cao, M., Sun, J., Wu, H., Zhao, P., et al. (2016). Response of MAPK
pathway to iron oxide nanoparticles in vitro treatment promotes osteogenic differentiation of
hBMSCs. Biomaterials, 86, 11–20.
136. Andre, R., Natalio, F., Humanes, M., Leppin, J., Heinze, K., Wever, R., et al. (2011). V2O5
Nanowires with an intrinsic peroxidase-like activity. Advanced Functional Materials, 21,
501–509.
137. Natalio, F., Andre, R., Hartog, A. F., Stoll, B., Jochum, K. P., Wever, R., et al. (2012).
Vanadium pentoxide nanoparticles mimic vanadium haloperoxidases and thwart biofilm
formation. Nature Nanotechnology, 7, 530–535.
138. Vernekar, A. A., Sinha, D., Srivastava, S., Paramasivam, P. U., D’Silva, P., & Mugesh, G.
(2014). An antioxidant nanozyme that uncovers the cytoprotective potential of vanadia
nanowires. Nature Communications, 5, 5301.
139. Han, L., Zeng, L. X., Wei, M. D., Li, C. M., & Liu, A. H. (2015). A V2O3-ordered
mesoporous carbon composite with novel peroxidase-like activity towards the glucose
colorimetric assay. Nanoscale, 7, 11678–11685.
140. Deng, H. M., Shen, W., Peng, Y. F., Chen, X. J., Yi, G. S., & Gao, Z. Q. (2012).
Nanoparticulate peroxidase/catalase mimetic and its application. Chemistry-A European
Journal, 18, 8906–8911.
141. Wu, M., Meng, S., Wang, Q., Si, W., Huang, W., & Dong, X. (2015). Nickel-cobalt oxide
decorated three-dimensional graphene as an enzyme mimic for glucose and calcium
detection. ACS Applied Materials & Interfaces, 7, 21089–21094.
142. Qu, K. G., Shi, P., Ren, J. S., & Qu, X. G. (2014). Nanocomposite incorporating V2O5
nanowires and gold nanoparticles for mimicking an enzyme cascade reaction and its
application in the detection of biomolecules. Chemistry-A European Journal, 20, 7501–7506.
143. Cha, S.-H., Hong, J., McGuffie, M., Yeom, B., VanEpps, J. S., & Kotov, N. A. (2015).
Shape-dependent biomimetic inhibition of enzyme by nanoparticles and their antibacterial
activity. ACS Nano, 9, 9097–9105.
144. Wan, Y., Qi, P., Zhang, D., Wu, J. J., & Wang, Y. (2012). Manganese oxide
nanowire-mediated enzyme-linked immunosorbent assay. Biosensors & Bioelectronics, 33,
69–74.
145. Zhang, X. D., & Huang, Y. M. (2015). Evaluation of the antioxidant activity of phenols and
tannic acid determination with Mn3O4 nano-octahedrons as an oxidase mimic. Analytical
Methods, 7, 8640–8646.
146. Jin, L. Y., Dong, Y. M., Wu, X. M., Cao, G. X., & Wang, G. L. (2015). Versatile and
amplified biosensing through enzymatic cascade reaction by coupling alkaline phosphatase
in situ generation of photoresponsive nanozyme. Analytical Chemistry, 87, 10429–10436.
147. Chen, W., Chen, J., Feng, Y. B., Hong, L., Chen, Q. Y., Wu, L. F., et al. (2012).
Peroxidase-like activity of water-soluble cupric oxide nanoparticles and its analytical
application for detection of hydrogen peroxide and glucose. Analyst, 137, 1706–1712.
148. Zhang, L. L., Han, L., Hu, P., Wang, L., & Dong, S. J. (2013). TiO2 nanotube arrays:
Intrinsic peroxidase mimetics. Chemical Communications, 49, 10480–10482.
149. Ragg, R., Natalio, F., Tahir, M. N., Janssen, H., Kashyap, A., Strand, D., et al. (2014).
Molybdenum trioxide nanoparticles with intrinsic sulfite oxidase activity. ACS Nano, 8,
5182–5189.
150. Dong, J. L., Song, L. N., Yin, J. J., He, W. W., Wu, Y. H., Gu, N., et al. (2014). Co3O4
nanoparticles with multi-enzyme activities and their application in immunohistochemical
assay. ACS Applied Materials & Interfaces, 6, 1959–1970.
151. Qu, K., Shi, P., Ren, J., & Qu, X. (2014). Nanocomposite incorporating V2O5 nanowires and
gold nanoparticles for mimicking an enzyme cascade reaction and its application in the
detection of biomolecules. Chemistry-A European Journal, 20, 7501–7506.
References 91

152. Mu, J. S., Zhang, L., Zhao, M., & Wang, Y. (2013). Co3O4 nanoparticles as an efficient
catalase mimic: Properties, mechanism and its electrocatalytic sensing application for
hydrogen peroxide. Journal of Molecular Catalysis A-Chemical, 378, 30–37.
153. Zhang, H., Pokhrel, S., Ji, Z., Meng, H., Wang, X., Lin, S., et al. (2014). PdO doping tunes
band-gap energy levels as well as oxidative stress responses to a Co3O4 p-type
semiconductor in cells and the lung. Journal of the American Chemical Society, 136,
6406–6420.
154. Mu, J. S., Zhang, L., Zhao, M., & Wang, Y. (2014). Catalase mimic property of Co3O4
nanomaterials with different morphology and its application as a calcium sensor. ACS
Applied Materials & Interfaces, 6, 7090–7098.
155. Chen, W., Chen, J., Liu, A. L., Wang, L. M., Li, G. W., & Lin, X. H. (2011). Peroxidase-like
activity of cupric oxide nanoparticle. ChemCatChem, 3, 1151–1154.
156. Hu, A. L., Liu, Y. H., Deng, H. H., Hong, G. L., Liu, A. L., Lin, X. H., et al. (2014).
Fluorescent hydrogen peroxide sensor based on cupric oxide nanoparticles and its application
for glucose and L-lactate detection. Biosensors & Bioelectronics, 61, 374–378.
157. Li, G.-W., Hong, L., Tong, M.-S., Deng, H.-H., Xia, X.-H., & Chen, W. (2015).
Determination of tannic acid based on luminol chemiluminescence catalyzed by cupric oxide
nanoparticles. Analytical Methods, 7, 1924–1928.
158. Wu, C. W., Harroun, S. G., Lien, C. W., Chang, H. T., Unnikrishnan, B., Lai, I. P. J., et al.
(2016). Self-templated formation of aptamer-functionalized copper oxide nanorods with
intrinsic peroxidase catalytic activity for protein and tumor cell detection. Sensors and
Actuators B-Chemical, 227, 100–107.
159. Applerot, G., Lellouche, J., Lipovsky, A., Nitzan, Y., Lubart, R., Gedanken, A., et al. (2012).
Understanding the antibacterial mechanism of cuo nanoparticles: Revealing the route of
induced oxidative stress. Small (Weinheim an der Bergstrasse, Germany), 8, 3326–3337.
160. Ragavan, K. V., & Rastogi, N. K. (2016). Graphene-copper oxide nanocomposite with
intrinsic peroxidase activity for enhancement of chemiluminescence signals and its
application for detection of bisphenol-A. Sensors and Actuators B-Chemical, 229, 570–580.
161. Li, G. L., Ma, P., Zhang, Y. F., Liu, X. L., Zhang, H., Xue, W. M., et al. (2016). Synthesis of
Cu2O nanowire mesocrystals using PTCDA as a modifier and their superior peroxidase-like
activity. Journal of Materials Science, 51, 3979–3988.
162. Liu, Y. J., Zhu, G. X., Bao, C. L., Yuan, A. H., & Shen, X. P. (2014). Intrinsic
peroxidase-like activity of porous CuO micro-/nanostructures with clean surface. Chinese
Journal of Chemistry, 32, 151–156.
163. Chen, W., Hong, L., Liu, A.-L., Liu, J.-Q., Lin, X.-H., & Xia, X.-H. (2012). Enhanced
chemiluminescence of the luminol-hydrogen peroxide system by colloidal cupric oxide
nanoparticles as peroxidase mimic. Talanta, 99, 643–648.
164. Hong, L., Liu, A.-L., Li, G.-W., Chen, W., & Lin, X.-H. (2013). Chemiluminescent
cholesterol sensor based on peroxidase-like activity of cupric oxide nanoparticles. Biosensors
& Bioelectronics, 43, 1–5.
Chapter 5
Other Nanomaterials for Nanozymes

Abstract The use of other nanomaterials beyond carbon-based nanomaterials,


metal-based nanomaterials, and metal oxide-based nanomaterials for mimicking
natural enzymes is discussed in this chapter. Prussian blue, metal-organic frame-
works, metal chalcogenides, metal hydroxides, etc., have been selected as repre-
sentative nanomaterials for mimicking peroxidase, superoxide dismutase, catalase,
etc. The catalytic mechanisms are also discussed if they have been elucidated.
Selected examples for in vitro biosensing, in vivo bioanalysis, and therapeutics are
discussed to highlight the broad applications of these nanozymes.

 
Keywords Nanozymes Artificial enzymes Integrated nanozymes Prussian 
  
blue Metal-organic frameworks Metal chalcogenides Metal hydroxides 
 
Cascade reactions Enzyme mimics Functional nanomaterials

As discussed in Chap. 1, more and more nanomaterials have been explored to


investigate their enzyme mimicking activities [1–60]. To highlight the ever-
growing interests in searching for new nanozymes, in this chapter selected exam-
ples of other nanomaterials for nanozymes are discussed.

5.1 Prussian Blue

Prussian blue, [Fe(III)Fe(II)(CN)6]−, based nanomaterials have been used to mimic


peroxidase, catalase, and SOD [4–6, 61]. In their initial report, Gu et al. found that
the Prussian blue coating could enhance the peroxidase mimicking activity of
γ-Fe2O3 nanoparticles [3]. Later, they showed that Prussian blue nanoparticles
exhibited catalase-like activity at neutral conditions (i.e., pH = 7.4) [4]. The
nanozyme was used for in vivo ultrasound and magnetic-resonance imaging of
overproduced H2O2 in diseased tissues. The nanozyme converted H2O2 into O2 gas
bubbles, which acted as the ultrasound contrast agents. Moreover, due to the
paramagnetic property of the formed O2 gas bubbles, they also acted as T1

© The Author(s) 2016 93


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_5
94 5 Other Nanomaterials for Nanozymes

Fig. 5.1 a Schematic showing intracellular behaviors of Prussian blue nanoparticles (PBNPs).
b Proposed mechanisms of the multiple enzyme-like activities of PBNPs based on standard redox
potentials of different compounds in the reaction systems. c Reactions involved. Prussian white
(PW, [Fe(II)Fe(II)(CN)6]2−), Berlin green (BG, {Fe(III)3[Fe(III)(CN)6]2[Fe(II)(CN)6]}2−), and
Prussian yellow (PY, [Fe(II)Fe(III)(CN)6]). Reprinted from Ref. [5], Copyright 2016, with
permission from American Chemical Society
5.1 Prussian Blue 95

magnetic-resonance contrast agents [4]. Using a lipopolysaccharide (LPS)-induced


liver inflammation model, the overproduced H2O2 in living mouse livers was
successfully imaged both by ultrasound and magnetic resonance methods.
Recently, Gu et al. [5] further confirmed that Prussian blue nanoparticles pos-
sessed peroxidase-like, catalase-like, and SOD-like activities (Fig. 5.1). Detailed
experimental results suggested that the peroxidase mimicking activity was domi-
nant at acidic conditions while the catalase mimicking activity was dominant at
high pH conditions. Interestingly, Prussian blue nanoparticles exhibited SOD-like
activities at different pH levels. Due to the mixed valence of Fe2+ and Fe3+, Prussian
blue can be reduced into Prussian white and oxidized into Berlin green or Prussian
yellow. As shown in Fig. 5.1b, the peroxidase-like activity was attributed to the
following mechanism: Prussian blue was first oxidized into Prussian yellow or
Berlin green by H2O2 at acidic pH (Eq. 5.6). Prussian yellow/Berlin green would
then oxidize TMB to into TMBox. Therefore, Prussian yellow/Berlin as the per-
oxidase mimic transferred electrons from TMB to H2O2 to achieve the catalytic
recycle (Eq. 5.7). Since H2O2 could act both as oxidizing and reducing agents, the
presence of the nanozyme would dismutate H2O2 into H2O and O2 via the mech-
anisms shown in Eqs. 5.8–5.11. Similarly, superoxide anion could be converted
into H2O2 and O2 via the mechanisms shown in Eqs. 5.12–5.16. They further
demonstrated that the nanozymes could alleviate inflammation by scavenging ROS
in vivo in lipopolysaccharide-treated mice model [5].

5.2 Metal-Organic Frameworks

Metal-organic frameworks (MOFs) themselves and MOFs loaded with other cata-
lysts have been used to mimic natural enzymes [7–18]. For instance, it has reported
that a well-known Cu2+ and benzene-1,3,5-tricarboxylate ligand-based MOF (i.e.,
HKUST-1) could mimic protease. Similar with natural trypsin, HKUST-1 could
catalyze the hydrolysis of bovine serum albumin (BSA) and casein. Moreover, in
comparison with natural trypsin, the MOF-based nanozyme had higher affinity
toward BSA [9].
For MOFs made from non-redox active metal ions, they could be imparted with
catalytic activities by encapsulating guest catalysts [7, 8, 15]. Usually, one catalyst
was usually encapsulated within MOFs for catalysis [15]. In biological systems,
however, multiple enzymes are usually compartmentalized within subcellular
organelles for cascade catalytic reactions. Such compartmentalization would lead to
an interesting “nanoscale proximity effects,” which could significantly enhance the
coupled reactions via several mechanisms. For example, it could improve the local
concentrations of catalysts and substrates, reduce the diffusion barrier, and stabilize
the catalysts and the unstable intermediates. Inspired by this, several strategies have
been developed to co-encapsulate two catalysts guests within various matrix [7, 8,
62–66].
96 5 Other Nanomaterials for Nanozymes

Wei and co-workers [7, 8] have developed a self-assembly strategy to fabricate


integrated nanozymes (INAzymes) by co-assembling multiple catalytic guests
within MOFs (Fig. 5.2). They first demonstrated that molecular catalysts (e.g.,
hemin) and natural enzymes (e.g., GOx) could be simultaneously confined with
ZIF-8, a MOF made from Zn2+ and 2-methylimidazole ligands, under biocom-
patible reaction conditions (Fig. 5.2a). Moreover, the obtained INAzyme of hemin
and GOx exhibited more than 600 % enhancement of the catalytic activity when
compared with the mixture of hemin@ZIF-8 and GOx@ZIF-8 (Fig. 5.2c). After
establishing that the INAzyme could be used for sensitive and selective detection of
glucose in vitro, they went on to construct an analytical platform by immobilizing
the INAzyme into the channel of a microfluidics chip. When further assisted with
microdialysis, the dynamic changes of brain glucose following ischemia and per-
fusion has been successfully monitored with the platform (Fig. 5.2d, e). Their
strategy was general and applicable to other combination of catalyst guests (such as
hemin/lactate oxidase and hemin/GOx/invertase) [8].

Fig. 5.2 Integrated nanozymes for monitoring the dynamic changes of brain glucose following
ischemia/reperfusion. a Schematic and TEM image of the integrated nanozymes (INAzymes).
b Reactions catalyzed by the INAzymes. c Normalized cascade catalytic activity of the INAzyme
(1), and the mixture of hemin@ZIF-8 and GOx@ZIF-8 (2), showing a more than 600 %
enhancement for the INAzyme when compared with the mixture of hemin@ZIF-8 and
GOx@ZIF-8. d Schematic illustration of the global cerebral ischemia. e Continuously monitoring
the dynamic changes of glucose level in the striatum of a living rat brain following global
ischemia/reperfusion with the INAzyme-based sensing platform. Adapted from Ref. [8], Copyright
2016, with permission from American Chemical Society
5.3 Metal Chalcogenides 97

5.3 Metal Chalcogenides

Numerous studies have showed that metal chalcogenides (such as CuS, MnSe, and
FeSe) could mimic peroxidase [19–34]. For example, Huang et al. prepared NiTe
nanowires via a hydrothermal method. The obtained NiTe nanowires showed
peroxidase mimicking activity, which were better than the commercial NiTe
powders. They further combined the nanozyme with GOx for sensitive and selective
detection of glucose [24].

5.4 Metal Hydroxides

Metal hydroxides, including layered double hydroxides, have been explored to


mimic peroxidase [35–38]. For example, Sun et al. reported the peroxidase-like
activity of CoFe layered double hydroxides and used them for colorimetric
detection of H2O2 and glucose [37]. Tan’s group fabricated an interesting Cu(OH)2
nanocages from amorphous Cu(OH)2 nanoparticles. Catalytic studies revealed that
the nanocages showed higher peroxidase mimicking activities even than that of
natural enzymes (Fig. 5.3) [38].

Fig. 5.3 Cu(OH)2 nanocages


as peroxidase mimics.
Reprinted from Ref. [38],
Copyright 2015, with
permission from American
Chemical Society
98 5 Other Nanomaterials for Nanozymes

Fig. 5.4 a SEM image of polypyrrole/hemin nanocomposites. b Polypyrrole/hemin nanocom-


posites as peroxidase mimics for glucose sensing. Adapted from Ref. [54], Copyright 2014, with
permission from American Chemical Society

5.5 Miscellaneous

Lots of other materials have been reported to mainly mimic peroxidase [39–60, 67].
For example, magnetic zirconium hexacyanoferrate (II) nanoparticles as peroxidase
mimics were used to label probe ssDNA for electrochemical DNA sensing [56].
Dong et al. [54] developed a facile way to prepare polypyrrole/hemin nanocom-
posites by chemical oxidative polymerization of pyrrole monomer with FeCl3 in the
presence of hemin (Fig. 5.4). The peroxidase-like activity of the nanocomposites
was confirmed by catalyzing the oxidation of TMB and other peroxidase substrates
with H2O2. When the peroxidase mimics were combined with GOx, the selective
and sensitive detection of glucose was achieved [54].

References

1. Wei, H., & Wang, E. K. (2013). Nanomaterials with enzyme-like characteristics (nanozymes):
Next-generation artificial enzymes. Chemical Society Reviews, 42, 6060–6093.
2. Wang, X. Y., Hu, Y. H., & Wei, H. (2016). Nanozymes in bionanotechnology: From sensing
to therapeutics and beyond. Inorganic Chemistry Frontiers, 3, 41–60.
3. Zhang, X.-Q., Gong, S.-W., Zhang, Y., Yang, T., Wang, C.-Y., & Gu, N. (2010). Prussian
blue modified iron oxide magnetic nanoparticles and their high peroxidase-like activity.
Journal of Materials Chemistry, 20, 5110–5116.
4. Yang, F., Hu, S., Zhang, Y., Cai, X., Huang, Y., Wang, F., et al. (2012). A hydrogen
peroxide-responsive O2 nanogenerator for ultrasound and magnetic-resonance dual modality
imaging. Advanced Materials, 24, 5205–5211.
5. Zhang, W., Hu, S. L., Yin, J. J., He, W. W., Lu, W., Ma, M., et al. (2016). Prussian blue
nanoparticles as multienzyme mimetics and reactive oxygen species scavengers. Journal of the
American Chemical Society, 138, 5860–5865.
References 99

6. Liu, T., Niu, X., Shi, L., Zhu, X., Zhao, H., & Lana, M. (2015). Electrocatalytic analysis of
superoxide anion radical using nitrogen-doped graphene supported Prussian Blue as a
biomimetic superoxide dismutase. Electrochimica Acta, 176, 1280–1287.
7. Cheng, H. J., Wei, H. (2014). Gordon Research Conference: Bioanalytical Sensors, Newport,
RI, USA, June 22–27, 2014.
8. Cheng, H. J., Zhang, L., He, J., Guo, W. J., Zhou, Z. Y., Zhang, X. J., et al. (2016). Integrated
nanozymes with nanoscale proximity for in vivo neurochemical monitoring in living brains.
Analytical Chemistry, 88, 5489–5497.
9. Li, B., Chen, D. M., Wang, J. Q., Yan, Z. Y., Jiang, L., Duan, D. L., et al. (2014). MOFzyme:
Intrinsic protease-like activity of Cu-MOF. Scientific Reports, 4, 6759.
10. Dong, W. F., Liu, X. D., Shi, W. B., & Huang, Y. M. (2015). Metal-organic framework
MIL-53(Fe): Facile microwave-assisted synthesis and use as a highly active peroxidase
mimetic for glucose biosensing. RSC Advances, 5, 17451–17457.
11. Ai, L. H., Li, L. L., Zhang, C. H., Fu, J., & Jiang, J. (2013). MIL-53(Fe): A metal-organic
framework with intrinsic peroxidase-like catalytic activity for colorimetric biosensing.
Chemistry-A European Journal, 19, 15105–15108.
12. Cui, F. J., Deng, Q. F., & Sun, L. (2015). Prussian blue modified metal-organic framework
MIL-101(Fe) with intrinsic peroxidase-like catalytic activity as a colorimetric biosensing
platform. RSC Advances, 5, 98215–98221.
13. Deleu, W. P. R., Rivero, G., Teixeira, R. F. A., Du Prez, F. E., & De Vos, D. E. (2015). Metal
organic frameworks encapsulated in photocleavable capsules for uv-light triggered catalysis.
Chemistry of Materials, 27, 5495–5502.
14. Xiong, Y. H., Chen, S. H., Ye, F. G., Su, L. J., Zhang, C., Shen, S. F., et al. (2015). Synthesis
of a mixed valence state Ce-MOF as an oxidase mimetic for the colorimetric detection of
biothiols. Chemical Communications, 51, 4635–4638.
15. Qin, F.-X., Jia, S.-Y., Wang, F.-F., Wu, S.-H., Song, J., & Liu, Y. (2013).
Hemin@metal-organic framework with peroxidase-like activity and its application to
glucose detection. Catalysis Science & Technology, 3, 2761–2768.
16. Liu, Y. L., Zhao, X. J., Yang, X. X., & Li, Y. F. (2013). A nanosized metal-organic framework
of Fe-MIL-88NH2 as a novel peroxidase mimic used for colorimetric detection of glucose.
Analyst, 138, 4526–4531.
17. Lu, J., Xiong, Y., Liao, C., & Ye, F. (2015). Colorimetric detection of uric acid in human urine
and serum based on peroxidase mimetic activity of MIL-53(Fe). Analytical Methods, 7,
9894–9899.
18. Chen, D. M., Li, B., Jiang, L., Duan, D. L., Li, Y. Z., Wang, J. Q., et al. (2015). Highly
efficient colorimetric detection of cancer cells utilizing Fe-MIL-101 with intrinsic
peroxidase-like catalytic activity over a broad pH range. RSC Advances, 5, 97910–97917.
19. He, W. W., Jia, H. M., Li, X. X., Lei, Y., Li, J., Zhao, H. X., et al. (2012). Understanding the
formation of CuS concave superstructures with peroxidase-like activity. Nanoscale, 4,
3501–3506.
20. Chen, Q., Chen, J., Gao, C. J., Zhang, M. L., Chen, J. Y., & Qiu, H. D. (2015).
Hemin-functionalized WS2 nanosheets as highly active peroxidase mimetics for label-free
colorimetric detection of H2O2 and glucose. Analyst, 140, 2857–2863.
21. Qiao, F., Wang, Z., Xu, K., & Ai, S. (2015). Double enzymatic cascade reactions within
FeSe-Pt@SiO2 nanospheres: Synthesis and application toward colorimetric biosensing of
H2O2 and glucose. Analyst, 140, 6684–6691.
22. Guan, J. F., Peng, J., & Jin, X. Y. (2015). Synthesis of copper sulfide nanorods as peroxidase
mimics for the colorimetric detection of hydrogen peroxide. Analytical Methods, 7,
5454–5461.
23. Lin, T. R., Zhong, L. S., Song, Z. P., Guo, L. Q., Wu, H. Y., Guo, Q. Q., et al. (2014). Visual
detection of blood glucose based on peroxidase-like activity of WS2 nanosheets. Biosensors &
Bioelectronics, 62, 302–307.
100 5 Other Nanomaterials for Nanozymes

24. Wan, L. J., Liu, J. H., & Huang, X. J. (2014). Novel magnetic nickel telluride nanowires
decorated with thorns: synthesis and their intrinsic peroxidase-like activity for detection of
glucose. Chemical Communications, 50, 13589–13591.
25. Yao, W. T., Zhu, H. Z., Li, W. G., Yao, H. B., Wu, Y. C., & Yu, S. H. (2013). Intrinsic
peroxidase catalytic activity of Fe7S8 nanowires templated from [Fe16S20]/diethylenetriamine
hybrid nanowires. ChemPlusChem, 78, 723–727.
26. Wang, Y., Zhang, D., & Xiang, Z. (2015). Synthesis of α-MnSe crystal as a robust peroxidase
mimic. Materials Research Bulletin, 67, 152–157.
27. Cai, Q., Lu, S. K., Liao, F., Li, Y. Q., Ma, S. Z., & Shao, M. W. (2014). Catalytic degradation
of dye molecules and in situ SERS monitoring by peroxidase-like Au/CuS composite.
Nanoscale, 6, 8117–8123.
28. Dalui, A., Pradhan, B., Thupakula, U., Khan, A. H., Kumar, G. S., Ghosh, T., et al. (2015).
Insight into the mechanism revealing the peroxidase mimetic catalytic activity of quaternary
CuZnFeS nanocrystals: Colorimetric biosensing of hydrogen peroxide and glucose.
Nanoscale, 7, 9062–9074.
29. Lin, T. R., Zhong, L. S., Guo, L. Q., Fu, F. F., & Chen, G. N. (2014). Seeing diabetes: Visual
detection of glucose based on the intrinsic peroxidase-like activity of MoS2 nanosheets.
Nanoscale, 6, 11856–11862.
30. Lu, X. F., Bian, X. J., Li, Z. C., Chao, D. M., & Wang, C. (2013). A facile strategy to decorate
Cu9S5 nanocrystals on polyaniline nanowires and their synergetic catalytic properties.
Scientific Reports, 3, 2955.
31. Dutta, A. K., Maji, S. K., Mondal, A., Karmakar, B., Biswas, P., & Adhikary, B. (2012). Iron
selenide thin film: Peroxidase-like behavior, glucose detection and amperometric sensing of
hydrogen peroxide. Sensors and Actuators B-Chemical, 173, 724–731.
32. Qiao, F. M., Chen, L. J., Li, X. N., Li, L. F., & Ai, S. Y. (2014). Peroxidase-like activity of
manganese selenide nanoparticles and its analytical application for visual detection of
hydrogen peroxide and glucose. Sensors and Actuators B-Chemical, 193, 255–262.
33. Dutta, A. K., Das, S., Samanta, S., Samanta, P. K., Adhikary, B., & Biswas, P. (2013). CuS
nanoparticles as a mimic peroxidase for colorimetric estimation of human blood glucose level.
Talanta, 107, 361–367.
34. Zhao, K., Gu, W., Zheng, S., Zhang, C., & Xian, Y. (2015). SDS-MoS2 nanoparticles as
highly-efficient peroxidase mimetics for colorimetric detection of H2O2 and glucose. Talanta,
141, 47–52.
35. Wang, Y. L., Chen, S. H., Ni, F., Gao, F., & Li, M. G. (2009). Peroxidase-like layered double
hydroxide nanoflakes for electrocatalytic reduction of H2O2. Electroanalysis, 21, 2125–2132.
36. Cui, L., Yin, H. S., Dong, J., Fan, H., Liu, T., Ju, P., et al. (2011). A mimic peroxidase
biosensor based on calcined layered double hydroxide for detection of H2O2. Biosensors &
Bioelectronics, 26, 3278–3283.
37. Zhang, Y. W., Tian, J. Q., Liu, S., Wang, L., Qin, X. Y., Lu, W. B., et al. (2012). Novel
application of CoFe layered double hydroxide nanoplates for colorimetric detection of H2O2
and glucose. Analyst, 137, 1325–1328.
38. Cai, R., Yang, D., Peng, S. J., Chen, X. G., Huang, Y., Liu, Y., et al. (2015). Single
nanoparticle to 3D supercage: Framing for an artificial enzyme system. Journal of the
American Chemical Society, 137, 13957–13963.
39. Zhang, X. D., He, S. H., Chen, Z. H., & Huang, Y. M. (2013). CoFe2O4 nanoparticles as
oxidase mimic-mediated chemiluminescence of aqueous luminol for sulfite in white wines.
Journal of Agricultural and Food Chemistry, 61, 840–847.
40. Su, L., Qin, W. J., Zhang, H. G., Rahman, Z. U., Ren, C. L., Ma, S. D., et al. (2015). The
peroxidase/catalase-like activities of MFe2O4 (M = Mg, Ni, Cu) MNPs and their application in
colorimetric biosensing of glucose. Biosensors & Bioelectronics, 63, 384–391.
41. Fan, Y. W., Shi, W. B., Zhang, X. D., & Huang, Y. M. (2014). Mesoporous material-based
manipulation of the enzyme-like activity of CoFe2O4 nanoparticles. Journal of Materials
Chemistry A, 2, 2482–2486.
References 101

42. Yang, W., Hao, J., Zhang, Z., Lu, B., Zhang, B., & Tang, J. (2014). CoxFe3-xO4 hierarchical
nanocubes as peroxidase mimetics and their applications in H2O2 and glucose detection. RSC
Advances, 4, 35500–35504.
43. Abdolmohammad-Zadeh, H., & Rahimpour, E. (2015). A novel chemosensor for Ag(I) ion
based on its inhibitory effect on the luminol-H2O2 chemiluminescence response improved by
CoFe2O4 nanoparticles. Sensors and Actuators B-Chemical, 209, 496–504.
44. Kim, Y. S., & Jurng, J. (2013). A simple colorimetric assay for the detection of metal ions
based on the peroxidase-like activity of magnetic nanoparticles. Sensors and Actuators
B-Chemical, 176, 253–257.
45. Shi, W., Zhang, X., He, S., Li, J., & Huang, Y. (2013). Fast screening of the
nanoparticles-based enzyme mimetics by a chemiluminescence method. Scientia Sinica
Chimica, 43, 1591–1598.
46. Liu, S., Wang, L., Zhai, J. F., Luo, Y. L., & Sun, X. P. (2011). Carboxyl functionalized
mesoporous polymer: A novel peroxidase-like catalyst for H2O2 detection. Analytical
Methods, 3, 1475–1477.
47. Zhang, R. Z., He, S. J., Zhang, C. M., & Chen, W. (2015). Three-dimensional Fe- and
N-incorporated carbon structures as peroxidase mimics for fluorescence detection of hydrogen
peroxide and glucose. Journal of Materials Chemistry B, 3, 4146–4154.
48. Singh, A., Patra, S., Lee, J. A., Park, K. H., & Yang, H. (2011). An artificial enzyme-based
assay: DNA detection using a peroxidase-like copper-creatinine complex. Biosensors &
Bioelectronics, 26, 4798–4803.
49. Chen, X., Zhou, X. D., & Hu, J. M. (2012). Pt-DNA complexes as peroxidase mimetics and
their applications in colorimetric detection of H2O2 and glucose. Analytical Methods, 4,
2183–2187.
50. Wang, J. J., Han, D. X., Wang, X. H., Qi, B., & Zhao, M. S. (2012). Polyoxometalates as
peroxidase mimetics and their applications in H2O2 and glucose detection. Biosensors &
Bioelectronics, 36, 18–21.
51. Li, Y., Li, X., Wong, Y.-S., Chen, T., Zhang, H., Liu, C., et al. (2011). The reversal of
cisplatin-induced nephrotoxicity by selenium nanoparticles functionalized with
11-mercapto-1-undecanol by inhibition of ROS-mediated apoptosis. Biomaterials, 32,
9068–9076.
52. Li, Y., Li, X., Zheng, W., Fan, C., Zhang, Y., & Chen, T. (2013). Functionalized selenium
nanoparticles with nephroprotective activity, the important roles of ROS-mediated signaling
pathways. Journal of Materials Chemistry B, 1, 6365–6372.
53. Wang, W., Jiang, X. P., & Chen, K. Z. (2012). CePO4: Tb, Gd hollow nanospheres as
peroxidase mimic and magnetic-fluorescent imaging agent. Chemical Communications, 48,
6839–6841.
54. Hu, P., Han, L., & Dong, S. (2014). A facile one-pot method to synthesize a
polypyrrole/hemin nanocomposite and its application in biosensor, dye removal, and
photothermal therapy. ACS Applied Materials & Interfaces, 6, 500–506.
55. Ryabov, A. D., Cerón-Camacho, R., Saavedra-Díaz, O., Denardo, M. A., Ghosh, A., Le
Lagadec, R., et al. (2012). TAML activator-based amperometric analytical devices as
alternatives to peroxidase biosensors. Analytical Chemistry, 84, 9096–9100.
56. Zhang, G.-Y., Deng, S.-Y., Cai, W.-R., Cosnier, S., Zhang, X.-J., & Shan, D. (2015).
Magnetic zirconium hexacyanoferrate(ii) nanoparticle as tracing tag for electrochemical dna
assay. Analytical Chemistry, 87, 9093–9100.
57. Tang, Y. R., Zhang, Y., Liu, R., Su, Y. Y., & Lu, Y. (2013). Application of NaYF4: Yb, Er
nanoparticles as peroxidase mimetics in uric acid detection. Chinese Journal of Analytical
Chemistry, 41, 330–336.
58. Sanchez-Sanchez, A., Arbe, A., Colmenero, J., & Pomposo, J. A. (2014). Metallo-folded
single-chain nanoparticles with catalytic selectivity. ACS Macro Letters, 3, 439–443.
59. Tao, Y., Ju, E. G., Ren, J. S., & Qu, X. G. (2014). Polypyrrole nanoparticles as promising
enzyme mimics for sensitive hydrogen peroxide detection. Chemical Communications, 50,
3030–3032.
102 5 Other Nanomaterials for Nanozymes

60. Liu, Y. J., Zhu, G. X., Yang, J., Yuan, A. H., & Shen, X. P. (2014). Peroxidase-like catalytic
activity of Ag3PO4 nanocrystals prepared by a colloidal route. PLoS ONE, 9, 0109158.
61. Zhang, X. Q., Gong, S. W., Zhang, Y., Yang, T., Wang, C. Y., & Gu, N. (2010). Prussian blue
modified iron oxide magnetic nanoparticles and their high peroxidase-like activity. Journal of
Materials Chemistry, 20, 5110–5116.
62. Huang, Y., Lin, Y., Ran, X., Ren, J., & Qu, X. (2016). Self-assembly and
compartmentalization of nanozymes in mesoporous silica-based nanoreactors. Chemistry—A
European Journal, 22, 5705–5711.
63. Kong, J., Yu, X., Hu, W., Hu, Q., Shui, S., Li, L., et al. (2015). A biomimetic enzyme
modified electrode for H2O2 highly sensitive detection. Analyst, 140, 7792–7798.
64. Lin, Y. H., Wu, L., Huang, Y. Y., Ren, J. S., & Qu, X. G. (2015). Positional assembly of
hemin and gold nanoparticles in graphene-mesoporous silica nanohybrids for tandem
catalysis. Chemical Science, 6, 1272–1276.
65. Liang, H., Jiang, S., Yuan, Q., Li, G., Wang, F., Zhang, Z., et al. (2016). Co-immobilization of
multiple enzymes by metal coordinated nucleotide hydrogel nanofibers: Improved stability and
an enzyme cascade for glucose detection. Nanoscale, 8, 6071–6078.
66. Lin, Y. H., Li, Z. H., Chen, Z. W., Ren, J. S., & Qu, X. G. (2013). Mesoporous
silica-encapsulated gold nanoparticles as artificial enzymes for self-activated cascade catalysis.
Biomaterials, 34, 2600–2610.
67. Huang, Y. Y., Ran, X., Lin, Y. H., Ren, J. S., & Qu, X. G. (2015). Self-assembly of an
organic-inorganic hybrid nanoflower as an efficient biomimetic catalyst for self-activated
tandem reactions. Chemical Communications, 51, 4386–4389.
Chapter 6
Challenges and Perspectives

Abstract The challenges and perspectives in the field of nanozymes are summa-
rized in this chapter, which if fully addressed, will lead to substantial breakthroughs
in the future.

 
Keywords Nanozymes Artificial enzymes Enzyme mimics Catalytic nano- 
  
materials Nanobiology Nanozymology Functional nanomaterials Biological 
 
catalysts Translational medicine Challenges and perspectives

As has been evidenced in the preceding chapters, highly active research interests
have been devoted to the field of nanozymes owing to their unique characteristics.
Despite the remarkable progress has been made in the field, the nanozyme research is
still in its infancy. Therefore, substantial breakthroughs are expected, which will lead
to the next wave of artificial enzymes for both fundamental science and practical
applications by overcoming the following as well as other challenges [1–4].
(1) Nanozymes with new catalytic properties beyond redox enzyme mimics
The currently developed nanozymes are mainly mimic redox enzymes (such as
peroxidase, oxidase, catalase, and superoxide dismutase) and hydrolytic enzymes
(such as nuclease, esterase) though other enzyme mimics have been reported [5, 6].
Given the fact that there are six major types of natural enzymes, future efforts
should be focused on designing new nanozymes that may mimic their
functionalities.
More, as Professor Dr. Jean-Marie Lehn stated: “the chemist finds inspiration in
the ingenuity of biological events and encouragement in the demonstration that
such high efficiencies, selectivities, and rates can indeed be attained. However,
chemistry is not limited to systems similar to those found in biology, but is free to
create unknown species and to invent novel processes,” researchers are therefore
encouraged to have great ambition to design new nanozymes beyond the natural
ones [7].

© The Author(s) 2016 103


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9_6
104 6 Challenges and Perspectives

(2) Rational design of nanozymes


The successful tackling of the above-mentioned challenge relies on our capa-
bility of rationally designing new nanozymes. To fulfill this goal, one should first
understand the nanozymes’ catalytic mechanisms both experimentally and theo-
retically [8–11]. Advanced characterization techniques, such as in situ (sub)atomic
resolution TEM imaging and synchrotron spectroscopies, are expected to provide
deeper insights. Recent progress in computational chemistry has allowed to design
functional proteins (including enzymes), which should provide invaluable guidance
to nanozyme design [12–14].
Despite the substantial progress in nanotechnology, it remains a great challenge
to prepare uniform (especially the atomically uniform) nanomaterials. On the other
hand, natural enzymes have well-defined amino acid sequences (or nucleic acid
sequences) as well as three-dimensional structures. Therefore, better synthetic
strategies are still needed to prepare highly efficient nanozymes with defined sizes
and structures even one may rationally design them in the future.
For practical applications, large-scale synthesis of nanozymes is also expected in
the near future. Industrial standards (and other necessary standards) should be
established for both nanozymes’ synthesis and characterization.
(3) Regulation of nanozymes’ activities
In biological systems, enzymes’ activities are highly regulated. Inspired by this
interesting phenomenon, numerous methods have been developed to tune the
nanozymes’ activities [1, 15–17]. In nature, an enzyme’s activity could be tuned by
regulating its expression and composition at genetic level and by controlling its
surrounding environment and its interaction with specific ligands. All of these
approaches should be exploited to regulate the nanozymes’ activities. For instance, in
a cell, multiple enzymes usually work together within confined compartments for
synergetic cascade reactions. Recent studies demonstrated that nanozymes could also
work cooperatively by co-assembling them together within confined spaces [15].
To obtain highly efficient nanozymes, the catalytic activities from both the core
and the surface coating should be exploited. For the nanozymes with catalytically
active core, the surface coating may shield their activities. Therefore, suitable
coatings should be adopted. On the other hand, some nanozymes’ cores only acts as
the supporting material. In this case, highly efficient nanozymes could be obtained
by exploring the synergistic activities of both the inorganic core and the surface
coatings. Also by introducing chiral cores or surface coatings, nanozymes with
chiral selectivities would be produced [18].
(4) Biosafety of nanozymes
Both the dynamic and final fates of nanozymes should be systematically studied
at different levels to address the potential toxicity concerns [19–24]. For example,
previous studies showed that the nanoparticles’ surface properties would affect their
interaction with cell membrane and their subcellular localizations [25–27].
Biodistribution studies suggested that the nanozymes without targeting molecules
6 Challenges and Perspectives 105

would accumulate mainly in liver, spleen, and lung [28, 29]. Other factors, like
nanoparticles’ aspect ratio, may also affect their biodistributions [30].
Encouragingly, several studies showed that a few regulatory agencies approved
that nanomaterials (such as Resovist (Ferucarbotran), i.e., the commercial available
superparamagnetic iron oxide nanoparticles) have been used as nanozymes [31].
However, for most of the currently developed nanozymes, both their acute and
long-term biosafety should be evaluated before they could be applied in clinics.
(5) Translational promise of nanozymes
As closely related with the biosafety issues discussed above, the potential
applications of the nanozymes in translational medicine and clinics remain largely
unexplored. Though the therapeutic effects of several nanozymes have been
investigated, much more efforts are needed to translate the encouraging lab (and
even pre-clinical) results into clinics to benefit patients and the society [32].
(6) Beyond the catalysis
Though it has suggested that redox active nanozymes showed promising neu-
roprotection and antioxidation effects via catalytic mechanisms, their applications
definitely could be extended to other areas [33]. For instance, they could be used to
modulate enzymes’ activities in vitro [34]. Overall, we believe that nanozymes will
be widely adopted not only as highly efficient biocatalysts but also as versatile tools
in nanobiology, an interfaced area of nanotechnology, biology, biomedicine, etc.
We and others have also speculated that nanozymes may even play a role in the
early stages of life [2, 35].

References

1. Wei, H., & Wang, E. K. (2013). Nanomaterials with enzyme-like characteristics (nanozymes):
Next-generation artificial enzymes. Chemical Society Reviews, 42, 6060–6093.
2. Wang, X. Y., Hu, Y. H., & Wei, H. (2016). Nanozymes in bionanotechnology: From sensing
to therapeutics and beyond. Inorganic Chemistry Frontiers, 3, 41–60.
3. Kotov, N. A. (2010). Inorganic nanoparticles as protein mimics. Science, 330, 188–189.
4. Bornscheuer, U. T., Huisman, G. W., Kazlauskas, R. J., Lutz, S., Moore, J. C., & Robins, K.
(2012). Engineering the third wave of biocatalysis. Nature, 485, 185–194.
5. Pieters, G., & Prins, L. J. (2012). Catalytic self-assembled monolayers on gold nanoparticles.
New Journal of Chemistry, 36, 1931–1939.
6. Fillon, Y., Verma, A., Ghosh, P., Ernenwein, D., Rotello, V. M., & Chmielewski, J. (2007).
Peptide ligation catalyzed by functionalized gold nanoparticles. Journal of the American
Chemical Society, 129, 6676–6677.
7. Lehn, J. M. (2006). Supramolecular Chemistry: Concepts and Perspectives. Wiley-VCH:
Weinheim.
8. Zhang, W., Hu, S. L., Yin, J. J., He, W. W., Lu, W., Ma, M., et al. (2016). Prussian blue
nanoparticles as multienzyme mimetics and reactive oxygen species scavengers. Journal of the
American Chemical Society, 5860–5865.
9. Shen, X., Liu, W., Gao, X., Lu, Z., Wu, X., & Gao, X. (2015). Mechanisms of oxidase and
superoxide dismutation-like activities of gold, silver, platinum, and palladium, and their
106 6 Challenges and Perspectives

alloys: A general way to the activation of molecular oxygen. Journal of the American
Chemical Society, 137, 15882–15891.
10. Zhao, R., Zhao, X., & Gao, X. (2015). Molecular-level insights into intrinsic peroxidase-like
activity of nanocarbon oxides. Chemistry-A European Journal, 21, 960–964.
11. Ali, S. S., Hardt, J. I., & Dugan, L. L. (2008). SOD Activity of carboxyfullerenes predicts their
neuroprotective efficacy: A structure-activity study. Nanomedicine-Nanotechnology Biology
and Medicine, 4, 283–294.
12. Gonen, S., DiMaio, F., Gonen, T., & Baker, D. (2015). Design of ordered two-dimensional
arrays mediated by noncovalent protein-protein interfaces. Science, 348, 1365–1368.
13. Samish, I., MacDermaid, C. M., Perez-Aguilar, J. M., & Saven, J. G. (2011). Theoretical and
computational protein design. Annual Review of Physical Chemistry, 62, 129–149.
14. Lu, Y., Yeung, N., Sieracki, N., & Marshall, N. M. (2009). Design of functional
metalloproteins. Nature, 460, 855–862.
15. Cheng, H. J., Zhang, L., He, J., Guo, W. J., Zhou, Z. Y., Zhang, X. J., et al. (2016). Integrated
nanozymes with nanoscale proximity for in vivo neurochemical monitoring in living brains.
Analytical Chemistry, 88, 5489–5497.
16. Tonga, G. Y., Jeong, Y., Duncan, B., Mizuhara, T., Mout, R., Das, R., et al. (2015).
Supramolecular regulation of bioorthogonal catalysis in cells using nanoparticle-embedded
transition metal catalysts. Nature Chemistry, 7, 597–603.
17. Song, Y., Xu, C., Wei, W., Ren, J., & Qu, X. (2011). Light regulation of peroxidase activity
by spiropyran functionalized carbon nanotubes used for label-free colorimetric detection of
lysozyme. Chemical Communications, 47, 9083–9085.
18. Chen, J. L.-Y., Pezzato, C., Scrimin, P., Prins, L. J. (2016). Chiral nanozymes–gold
nanoparticle-based transphosphorylation catalysts capable of enantiomeric discrimination.
Chemistry—A European Journal, 22, 7028–7032.
19. Xia, T., Kovochich, M., Liong, M., Madler, L., Gilbert, B., Shi, H. B., et al. (2008).
Comparison of the mechanism of toxicity of zinc oxide and cerium oxide nanoparticles based
on dissolution and oxidative stress properties. ACS Nano, 2, 2121–2134.
20. Minarchick, V. C., Stapleton, P. A., Porter, D. W., Wolfarth, M. G., Ciftyuerek, E., Barger,
M., et al. (2013). Pulmonary cerium dioxide nanoparticle exposure differentially impairs
coronary and mesenteric arteriolar reactivity. Cardiovascular Toxicology, 13, 323–337.
21. An, Q., Sun, C. Y., Li, D., Xu, K., Guo, J., & Wang, C. C. (2013). Peroxidase-like activity of
Fe3O4@carbon nanoparticles enhances ascorbic acid-induced oxidative stress and selective
damage to PC-3 prostate cancer cells. ACS Applied Materials & Interfaces, 5, 13248–13257.
22. See, V., Free, P., Cesbron, Y., Nativo, P., Shaheen, U., Rigden, D. J., et al. (2009).
Cathepsin L digestion of nanobioconjugates upon endocytosis. ACS Nano, 3, 2461–2468.
23. Wang, L. J., Min, Y., Xu, D. D., Yu, F. J., Zhou, W. Z., & Cuschieri, A. (2014). Membrane
lipid peroxidation by the peroxidase-like activity of magnetite nanoparticles. Chemical
Communications, 50, 11147–11150.
24. Szymanski, C. J., Munusamy, P., Mihai, C., Xie, Y., Hu, D., Gilles, M. K., et al. (2015). Shifts
in oxidation states of cerium oxide nanoparticles detected inside intact hydrated cells and
organelles. Biomaterials, 62, 147–154.
25. Vincent, A., Babu, S., Heckert, E., Dowding, J., Hirst, S. M., Inerbaev, T. M., et al. (2009).
Protonated nanoparticle surface governing ligand tethering and cellular targeting. ACS Nano,
3, 1203–1211.
26. Verma, A., Uzun, O., Hu, Y., Hu, Y., Han, H.-S., Watson, N., et al. (2008).
Surface-structure-regulated cell-membrane penetration by monolayer-protected
nanoparticles. Nature Materials, 7, 588–595.
27. Dowding, J. M., Das, S., Kumar, A., Dosani, T., McCormack, R., Gupta, A., et al. (2013).
Cellular interaction and toxicity depend on physicochemical properties and surface
modification of redox-active nanomaterials. ACS Nano, 7, 4855–4868.
28. Zhuang, J., Fan, K. L., Gao, L. Z., Lu, D., Feng, J., Yang, D. L., et al. (2012). Ex vivo
detection of iron oxide magnetic nanoparticles in mice using their intrinsic
peroxidase-mimicking activity. Molecular Pharmaceutics, 9, 1983–1989.
References 107

29. Rojas, S., Domingo Gispert, J., Abad, S., Buaki-Sogo, M., Garcia, H., & Raul Herance,
J. (2012). In vivo biodistribution of amino-functionalized ceria nanoparticles in rats using
positron emission tomography. Molecular Pharmaceutics, 9, 3543–3550.
30. Lin, S., Wang, X., Ji, Z., Chang, C. H., Dong, Y., Meng, H., et al. (2014). Aspect ratio plays a
role in the hazard potential of CeO2 nanoparticles in mouse lung and zebrafish gastrointestinal
tract. ACS Nano, 8, 4450–4464.
31. Huang, D. M., Hsiao, J. K., Chen, Y. C., Chien, L. Y., Yao, M., Chen, Y. K., et al. (2009). The
promotion of human mesenchymal stem cell proliferation by superparamagnetic iron oxide
nanoparticles. Biomaterials, 30, 3645–3651.
32. Dugan, L. L., Turetsky, D. M., Du, C., Lobner, D., Wheeler, M., Almli, C. R., et al. (1997).
Carboxyfullerenes as neuroprotective agents. Proceedings of the National Academy of
Sciences of the United States of America, 94, 9434–9439.
33. Dugan, L. L., Tian, L., Quick, K. L., Hardt, J. I., Karimi, M., Brown, C., et al. (2014).
Carboxyfullerene neuroprotection postinjury in parkinsonian nonhuman primates. Annals of
Neurology, 76, 393–402.
34. Huang, P.-J. J., Pautler, R., Shanmugaraj, J., Labbe, G., & Liu, J. (2015). Inhibiting the VIM-2
Metallo-β-Lactamase by graphene oxide and carbon nanotubes. ACS Applied Materials &
Interfaces, 7, 9898–9903.
35. Prins, L. J. (2015). Emergence of complex chemistry on an organic monolayer. Accounts of
Chemical Research, 48, 1920–1928.
Appendix

See Tables A.1, A.2 and A.3.

Table A.1 H2O2 detection with peroxidase mimics


Nanozymes Meth Linear range LOD Comments Ref.
Fe3O4 MNPs Color 5–100 μM 3 μM Substrate: ABTS [1]
Fe3O4 MNPs Color 0.5–150.0 μM 0.25 μM Substrate: DPD [2]
H2O2 in rainwater,
honey, and milk was
tested
Fe3O4 MNPs Color 1–100 μM 0.5 μM Substrate: TMB [3]
Fe3O4 was
encapsulated in
mesoporous silica
Fe3O4 graphene Color 1–50 μM 0.32 μM Substrate: TMB [4]
oxide composites
Fe-substituted SBA– Color 0.4–15 μM 0.2 μM Substrate: TMB [5]
15 microparticles
Iron phosphate Color 10–50 μM 10 nM Substrate: TMB [6]
microflowers
[Fe(III) Color 0.1–5 mM 10 μM Substrate: TMB [7]
(biuret-amide)] on
mesoporous silica
FeTe nanorods Color 0.1–5 μM 55 nM Substrate: ABTS [8]
Fe(III)-based Color 1–50 μM 0.4 μM Substrate: TMB [9]
coordination
polymer
Fe3O4 Color 5–80 μM 1.07 μM Substrate: TMB [10]
nanocomposites Fe3O4 was
functionalized by
5,10,15,20-Tetrakis
(4-carboxyphenyl)-
porphyrin
(continued)

© The Author(s) 2016 109


X. Wang et al., Nanozymes: Next Wave of Artificial Enzymes,
SpringerBriefs in Molecular Science, DOI 10.1007/978-3-662-53068-9
110 Appendix

Table A.1 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Protein-Fe3O4 and Color 0.5–200 μM 0.2 μM Substrate: TMB [11]
glucose oxidase
nanocomposites
GOx/Fe3O4/GO Color 0.1–100 μM 0.04 μM Substrate: DPD [12]
magnetic
nanocomposite
Iron(III) hydrogen Color 57.4–525.8 μM 1 μM Substrate: TMB [13]
phosphate hydrate
crystals
MIL-53(Fe) Color 0.95–19 μM 0.13 μM Substrate: TMB [14]
MIL-53(Fe): a metal–
organic framework
CuO NPs Color 0.01–1mM N/A Substrate: 4-AAP and [15]
phenol
AuNPs Color 18–1100 μM 4 μM Substrate: TMB [16]
Cysteamine was the
ligand for AuNPs
AuNC@BSA Color 0.5–20 μM 20 nM Substrate: TMB [17]
Au@Pt core/shell Color 45–1000 μM 45 μM Substrate: OPD [18]
nanorods
Nickel telluride Color 0.1–0.5 μM 25 nM Substrate: ABTS [19]
nanowires
Graphene oxide Color 0.05–100 μM 50 nM Substrate: TMB [20]
Hemin–graphene Color 0.05–500 μM 20 nM Substrate: TMB [21]
hybrid nanosheets
Carbon nanodots Color 1–100 μM 0.2 μM Substrate: TMB [22]
Carbon nitride dots Color 1–100 μM 0.4 μM Substrate: TMB [23]
Tungsten carbide Color 0.2–80 μM 60 nM Substrate: TMB [24]
nanorods
CoFe LDH Color 1–20 μM 0.4 μM Substrate: TMB [25]
nanoplates
CoxFe3–xO4 Color. 1–60 μM 0.36 μM Substrate: TMB [26]
nanocubes
Porphyrin Color 1–75 μM 0.4 μM Substrate: TMB [27]
functionalized
Co3O4
nanostructures
Carboxyl Color 1–8 μM 0.4 μM Substrate: TMB [28]
functionalized
mesoporous polymer
PtPd nanodendrites Color 0.5–150 μM 0.1 μM Substrate: TMB [29]
on graphene
nanosheets
(PtPdNDs/GNs)
(continued)
Appendix 111

Table A.1 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Pt-DNA complexes Color 0.979–17.6 mM 0.392 mM Substrate: TMB [30]
3.92 μM was detected
with PVDF
membrane
Manganese selenide Color 0.17–10 μM 0.085 μM Substrate: TMB [31]
nanoparticles
Prussian blue Color 0.05–50 μM 0.031 μM Substrate: ABTS [32]
nanoparticles
MWCNTs-Prussian Color 1 μM–1.5 mM 100 nM Substrate: TMB [33]
blue nanoparticles Carbon nanotubes
were filled with
Prussian blue
nanoparticles
Polypyrrole Color 5–100 μM Substrate: TMB [34]
nanoparticles PPy has been
successfully
employed to
quantitatively monitor
the H2O2 generated
by macrophages
Polyoxometalate Color 1–20 μM 0.4 μM Substrate: TMB [35]
Polyoxometalate Color 0.134–67 μM 0.134 μM Substrate: TMB [36]
Fe3O4 MNPs Fluor 10–200 nM 5.8 nM Substrate: Rhodamine [37]
B
Fluorescence of
Rhodamine B was
quenched
BiFeO3 NPs Fluor 20 nM–20 μM 4.5 nM Substrate: BA [38]
Oxidation of BA gave
fluorescence
H2O2 in rainwater
was tested
Fe3O4 MNPs Fluor 0.18–900 μM 0.18 μM Fluorescence of [39]
CdTe QD was
quenched
Fe3O4 MNPs Fluor 0.04–8 μM 0.008 μM Substrate: BA [40]
Oxidation of BA gave
fluorescence
cupric oxide Fluor 5–200 μM 0.34 μM Substrate: terephthalic [41]
nanoparticles acid
Terephthalic acid was
oxidized by hydroxyl
radical to form a
highly fluorescent
product
(continued)
112 Appendix

Table A.1 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Fe(III)–TAML CL 0.06–1 μM 0.05 μM [42]
activator
CoFe2O4 NPs CL 0.1–4 μM 0.02 μM CoFe2O4 NPs form [43]
complexes with
beta-CD
CoFe2O4 NPs CL 0.1–10 μM 10 nM H2O2 in natural water [44]
was tested
CoFe2O4 NPs with CL 1 nM–4 μM 0.5 nM CoFe2O4 NPs was [45]
chitosan coating coated with chitosan
H2O2 in natural water
was tested
Fe3O4 MNPs E-chem 4.2–800 μM 1.4 μM [46]
Fe3O4 E-chem 1.2–3500 μM 1.2 μM H2O2 in disinfected [47]
microspheres-AgNP FBS samples was
hybrids tested
Fe3O4 MNPs E-chem 0–16 nM 1.6 nM Fe3O4 was loaded on [48]
CNT
Fe3O4 MNPs E-chem 1–10 mM N/A Fe3O4 was entrapped [49]
in mesoporous carbon
foam, and the
composite was used
to construct a carbon
paste electrode
Not a linear response
Fe3O4 MNPs E-chem 20–6250 μM 2.5 μM Fe3O4 MNPs and [50]
PDDA–graphene
formed multilayer via
layer-by-layer
assembly
H2O2 in toothpaste
was tested
Fe3O4 E-chem 1–700 μM 1 μM H2O2 in Walgreens [51]
nanofilms on TiN antiseptic/oral
substrate debriding agent, Crest
whitening mouthwash
solution, Diet coke,
and Gatorade was
tested
Fe3O4 MNPs E-chem 0.2–2 mM 0.01 mM [52]
Fe3O4 MNPs E-chem 0.1–6 mM 3.2 μM Fe3O4 was on reduced [53]
graphene oxide
Fe2O3 NPs E-chem 20–140 μM 11 μM [54]
Fe2O3 NPs E-chem 20–300 μM 7 μM Fe2O3 was modified [54]
with Prussian blue.
Iron oxide NPs/CNT E-chem 0.099–6.54 mM 53.6 μM [55]
(continued)
Appendix 113

Table A.1 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Fe3O4/self-reduced E-chem 0.001–20 mM 0.17 μM Extracellular H2O2 [56]
graphene released from HeLa
nanocomposites cells stimulated by
CdTe quantum dots
(QDs) was established
by this approach
FeS nanosheet E-chem 0.5–150 μM 92 nM [57]
FeS needle E-chem 5–140 μM 4.3 μM [58]
FeSe NPs E-chem 5–100 μM 3.0 μM [58]
FeS E-chem 10–130 μM 4.03 μM [59]
Co3O4 NPs E-chem 0.05–25 mM 0.01 mM [60]
Hemin–graphene E-chem 0.5–400 μM 0.2 μM [21]
hybrid nanosheets
Layered double E-chem 1–240 μM 0.3 μM [61]
hydroxide–hemin
nanocomposite
Helical CNT E-chem 0.5–115 μM 0.12 μM [62]
LDH nanoflakes E-chem 12–254 μM 2.3 μM [63]
Calcined LDH E-chem 1–100 μM 0.5 μM [64]
CdS E-chem 1–1900 μM 0.28 μM [65]

Table A.2 Targets detection combining oxidases and peroxidase mimics


Nanozymes Meth Linear range LOD Comments Ref.
Glucose
Fe3O4 MNPs Color 50–1000 μM 30 μM Substrate: ABTS [1]
Selectivity against
sugars: fructose,
lactose, and maltose
Fe3O4 MNPs with Color 39–100 μM 30 μM Substrate: ABTS [66]
PDDA coating GOx was
electrostatically
assembled onto the
Fe3O4@PDDA
Glucose in serum
samples was tested
Compared with
glucometer
Selectivity against
sugars: galactose,
lactose, mannose,
maltose, arabinose,
cellobiose, raffinose,
and xylose
(continued)
114 Appendix

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Fe3O4 MNPs Color 30–1000 μM 3 μM Substrate: TMB [3]
Fe3O4 was
encapsulated in
mesoporous silica with
GOx
Showing the recycle
capability
Comparison between
free MNPs versus
encapsulated MNPs
Fe3O4 GO Color 2–200 μM 0.74 μM Substrate: TMB [4]
composites Glucose in urine was
tested
Fe3O4 Color 5–25 μM 2.21 μM Substrate: TMB [10]
nanocomposites Fe3O4 was
functionalized by
5,10,15,20-Tetrakis
(4-carboxyphenyl)-
porphyrin
Protein-Fe3O4 and Color 3–1000 μM 1.0 μM Substrate: TMB [11]
glucose oxidase
nanocomposites
γ-Fe2O3 Color 1–80 μM 0.21 μM Substrate: TMB [67]
nanoparticles Glucose in blood and
urine was tested
GOx/Fe3O4/GO Color 0.5–600 μM 0.2 μM Substrate: DPD [12]
magnetic
nanocomposite
Graphite-like carbon Color 5–100 μM 0.1 μM Substrate: TMB [68]
nitrides Glucose in serum was
tested
Iron oxide NPs Color 31.2–250 μM 8.5 μM Substrate: ABTS [69]
Iron oxide NPs was
coated with glycine
More robust than HRP
towards NaN3
inhibition
Iron oxide NPs Color 31.2–250 μM 15.8 μM Substrate: ABTS [69]
Iron oxide NPs was
coated with heparin
More robust than HRP
towards NaN3
inhibition
Iron oxide NPs Color 0.12–4 μM 0.5 μM Substrate: ABTS [70]
Iron oxide NPs was
coated with APTES
and MPTES
(continued)
Appendix 115

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
ZnFe2O4 Color 1.25–18.75 μM 0.3 μM Substrate: TMB [71]
Glucose in urine was
tested
[Fe(III) Color 20–300 μM 10 μM Substrate: TMB [7]
(biuret-amide)] on Glucose in mice blood
mesoporous silica plasma was tested
FeTe nanorods Color 1–100 μM 0.38 μM Substrate: ABTS [8]
Glucose in spiked
blood was tested
Fe(III)-based Color 2–20 μM 1 μM Substrate: TMB [9]
coordination polymer Glucose in serum was
tested
Mesoporous Fe2O3– Color 0.5–10 μM 0.5 μM Substrate: TMB [72]
graphene Glucose in serum was
nanostructures tested
CuO NPs Color 0.1–8 mM N/A Substrate: 4-AAP and [15]
phenol
V2O5 nanowires and Color 0–10 μM 0.5 μM Substrate: ABTS [73]
Gold nanoparticles
nanocomposite
AuNPs Color 2.0–200 μM 0.5 μM Substrate: TMB [16]
Cysteamine was the
ligand for AuNPs
Au@Pt core/shell Color 45–400 μM 45 μM Substrate: OPD [18]
nanorods
Nickel telluride Color 1–50 μM 0.42 μM Substrate: ABTS [19]
nanowires
Manganese selenide Color 8–50 μM 1.6 μM Substrate: TMB [31]
nanoparticles
Graphene oxide Color 1–20 μM 1 μM Substrate: TMB [20]
Glucose in blood and
fruit juice was tested
Graphene oxide Color 2.5–5 mM 0.5 μM Substrate: TMB [74]
Graphene oxide was
functionalized by
chitosan
Hemin–graphene Color 0.05–500 μM 30 nM Substrate: TMB [21]
hybrid nanosheets
Carbon nanodots Color 1–500 μM 1 μM Substrate: TMB [22]
Glucose in serum was
tested
Carbon nitride dots Color 1–5 μM 0.5 μM Substrate: TMB [23]
(continued)
116 Appendix

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
MWCNTs–Prussian Color 1 μM–1 mM 200 nM Substrate: TMB [33]
blue nanoparticles Carbon nanotubes
were filled with
Prussian blue
nanoparticles
CoFe LDH Color 1–10 mM 0.6 μM Substrate: TMB [25]
nanoplates
CoxFe3-xO4 Color 8–90 μM 2.47 μM Substrate: TMB [26]
nanocubes
MoS2 Color 5–150 μM 1.2 μM Substrate: TMB [75]
nanosheets Glucose in serum was
tested
Tungsten disulfide Color 5–300 μM 2.9 μM Substrate: TMB [76]
nanosheets Glucose in serum of
normal persons and
diabetes persons was
tested
Prussian blue Color 0.1–50 μM 0.03 μM Substrate: ABTS [32]
nanoparticles
Fe3O4 MNPs Fluor 1.6–160 μM 1.0 μM Fluorescence of [39]
CdTe QD was
quenched
Glucose in serum was
tested
Fe3O4 MNPs Fluor 0.05–10 μM 0.025 μM Substrate: benzoic acid [40]
Oxidation of BA gave
fluorescence
Glucose in serum was
tested
Fe3O4 MNPs with Fluor 3–9 μM 3 μM GOx was [77]
PDDA coating electrostatically
assembled onto the
Fe3O4@PDDA
Oxidation of AU gave
fluorescence
Glucose in serum was
tested
Selectivity against
sugars: arabinose,
cellobiose, galactose,
lactose, maltose,
raffinose, and xylose
BiFeO3 NPs Fluor 1–100 μM 0.5 μM Oxidation of BA gave [38]
fluorescence
Glucose in serum was
tested
CoFe2O4 NPs CL 0.1–10 μM 0.024 μM Other sugars [44]
(continued)
Appendix 117

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
CoFe2O4 NPs CL 0.05–10 μM 10 nM CoFe2O4 NPs were [45]
coated with chitosan
Glucose in serum was
tested
Hemin–graphene E-chem 0.5–400 μM 0.3 μM [21]
hybrid nanosheets
Fe3O4 MNPs E-chem 6–2200 μM 6 μM Glucose in serum was [78]
tested
Compared with
clinical analyzer
Nafion for high
selectivity against AA,
UA, sucrose, and
lactose
Fe3O4 MNPs E-chem 0.5–10 mM 0.2 mM Fe3O4 was [49]
encapsulated in
mesoporous carbon
with GOx, and the
composite was used to
construct a carbon
paste electrode
Comparison between
free MNPs vs
encapsulated MNPs
Fe3O4–enzyme– E-chem 0.5 μM– 0.3 μM Glucose in serum was [79]
polypyrrole 34 mM tested
nanoparticles
Ascorbic acid
MIL-53(Fe) Color 28.6–190.5 μM 15 μM Substrate: TMB [14]
MIL-53(Fe): A Metal–
Organic Framework
Dopamine
CoxFe3-xO4 Color 0.6–8 μM 0.13 μM Substrate: TMB [80]
nanoparticles Dopamine in serum
was tested
Thrombin
Ag/Pt bimetallic Color 1–50 nM 2.6 nM Ag/Pt bimetallic [81]
nanoclusters nanoclusters was
produced through a
DNA-templated
method
Glutathione
Fe-MIL-88NH2 Color 1–100 μM 0.45 μM Substrate: TMB [82]
MOF
Cysteine
Fe-MIL-88NH2 Color 1–80 μM 0.39 μM Substrate: TMB [82]
MOF
(continued)
118 Appendix

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Homocysteine
Fe-MIL-88NH2 Color 1–80 μM 0.40 μM Substrate: TMB [82]
MOF
Choline
Fe3O4 Fluor 20–100 μM 20 μM Choline oxidase was [77]
MNPs with PDDA electrostatically
coating assembled onto the
Fe3O4@PDDA
Oxidation of AU gave
fluorescence
Fe3O4 MNPs E-chem 1 nM–10 mM 0.1 nM Fe3O4 and choline [83]
(log) oxidase were
immobilized together
on electrode
Selectivity against AA
and UA
Platinum Color 6–400 μM 2.5 μM Substrate: N-ethyl-N- [84]
nanoparticles (3-sulfopropyl)-
3-methylaniline
sodium salt and
4-amino-antipyrine
Acetylcholine
Fe3O4 Color 100 nM– 39 nM Substrate: TMB [85]
nanospheres/reduced 10 mM
graphene oxide
Platinum Color 10–200 μM 2.84 μM Substrate: N-ethyl-N- [84]
nanoparticles (3-sulfopropyl)-3-
methylaniline sodium
salt and
4-amino-antipyrine
Glutathione
Carbon nanodots Color 0–7 μM 0.3 μM Substrate: TMB [86]
Cholesterol
Fe3O4 MNPs Color 10–250 μM 5 μM Substrate: TMB [3]
Fe3O4 was
encapsulated in
mesoporous silica with
cholesterol oxidase
Showing the recycle
capability
Comparison between
free MNPs versus
encapsulated MNPs
(continued)
Appendix 119

Table A.2 (continued)


Nanozymes Meth Linear range LOD Comments Ref.
Au@Pt core/shell Color 30–300 μM 30 μM Substrate: OPD [18]
nanorods
Galactose
Fe3O4 MNPs Color 10–200 mg/L 5 mg/L Substrate: ABTS [87]
Galactose in dried
blood samples from
normal persons and
patients was tested
Plates were used for
sensing
Fe3O4 MNPs with Fluor 2–80 μM 2 μM Galactose oxidase was [77]
PDDA coating electrostatically
assembled onto the
Fe3O4@PDDA
Oxidation of AU gave
fluorescence
Melamine
Bare gold Color 1–800 nM 0.2 nM Substrate: TMB [88]
nanoparticles
Kanamycin
Gold nanoparticles Color 1–100 nM 4.52 nM Substrate: TMB [89]
Gold nanoparticles
were modified by
kanamycin aptamer
Xanthine
AuNC@BSA Color 1–200 μM 0.5 μM Substrate: TMB [17]
Xanthine in serum and
urine samples was
tested
Mercury(II)
Ag nanoparticles Color 0.5–800 nM 0.125 nM Substrate: TMB [90]
Mercury(II) in blood
and wastewater was
tested
carbon nanodots Color 0–0.46 μM 23 nM Substrate: TMB [91]
Platinum Color 0.01–4 nM 8.5 pM Substrate: TMB [92]
Nanoparticle
Calcium ion
Co3O4 Nanomaterials E-chem 0.1–1 mM 4 μM The calcium ion in a [93]
milk sample was
tested
120 Appendix

Table A.3 Nanozyme as peroxidase mimics for immunoassay


Nanozyme Target Format Comments Ref.
Fe3O4 NPs with PreS1 Antigen–down [94]
dextran coating immunoassay
TnI Capture-detection
sandwich
immunoassay
Fe3O4 NPs with Mouse IgG Antigen-down [95]
chitosan coating immunoassay
CEA Capture-detection
sandwich
immunoassay
CEA Sandwich
immunoassay
Fe2O3 NPs with IgG Antigen-down [96]
Prussian blue immunoassay
coating
Ferric nanocore Avidin antigen-down Avidin-biotin interaction [97]
residing in immunoassay
ferritin Nitrated Sandwich
human immunoassay
ceruloplasmin
Fe(1–x)MnxFe2O4 Mouse IgG Antigen-down Both direct and indirect assay [98]
NPs with immunoassay
PMIDA coating
MnFe2O4 NPs Sticholysin II Antigen-down [99]
with citric acid immunoassay
coating
Fe-TAML human IgG Antigen-down Fe-TAML was encapsulated [100]
immunoassay Inside Mesoporous Silica
Nanoparticles
Co3O4 vascular Antigen-down [101]
nanoparticles endothelial immunoassay
growth factor
Platinum cytokeratin Sandwich [102]
nanoparticles 19 fragments immunoassay
Platinum folate Antigen-down [103]
Nanoparticles on receptors immunoassay
Graphene Oxide
Gold respiratory Sandwich The peroxidase-like activity [104]
nanoparticles– syncytial immunoassay of gold
graphene oxide virus Nanoparticles–graphene
hybrids oxide hybrids could be
enhanced by mercury(II)
Rod-shaped human IgG Antigen-down The detection limit can be as [105]
Au@PtCu immunoassay low as 90 pg/mL
(continued)
Appendix 121

Table A.3 (continued)


Nanozyme Target Format Comments Ref.
Au@Pt nanorods mouse IL-2 Sandwich [106]
with PSS coating immunoassay
Graphene oxide PSA Sandwich Clinical samples were tested [107]
immunoassay

References
1. Wei, H., & Wang, E. (2008). Fe3O4 magnetic nanoparticles as peroxidase mimetics and
their applications in H2O2 and glucose detection. Analytical Chemistry, 80, 2250–2254.
2. Chang, Q., Deng, K. J., Zhu, L. H., Jiang, G. D., Yu, C., & Tang, H. Q. (2009).
Determination of hydrogen peroxide with the aid of peroxidase-like Fe3O4 magnetic
nanoparticles as the catalyst. Microchimica Acta, 165, 299–305.
3. Kim, M. I., Shim, J., Li, T., Lee, J., & Park, H. G. (2011). Fabrication of nanoporous
nanocomposites entrapping Fe3O4 magnetic nanoparticles and oxidases for colorimetric
biosensing. Chemistry-A European Journal, 17, 10700–10707.
4. Dong, Y. L., Zhang, H. G., Rahman, Z. U., Su, L., Chen, X. J., Hu, J., et al. (2012).
Graphene oxide-Fe3O4 magnetic nanocomposites with peroxidase-like activity for colori-
metric detection of glucose. Nanoscale, 4, 3969–3976.
5. Liu, S., Tian, J. Q., Wang, L., Luo, Y. L., Chang, G. H., & Sun, X. P. (2011).
Iron-substituted SBA-15 microparticles: A peroxidase-like catalyst for H2O2 detection.
Analyst, 136, 4894–4897.
6. Wang, S., Chen, W., Liu, A. L., Hong, L., Deng, H. H., & Lin, X. H. (2012). Comparison
of the peroxidase-like activity of unmodified, amino-modified, and citrate-capped gold
nanoparticles. ChemPhysChem, 13, 1199–1204.
7. Malvi, B., Panda, C., & Dhar, B. B. (2012). Sen Gupta, S. One pot glucose detection by
[Fe-III(biuret-amide)] immobilized on mesoporous silica nanoparticles: An efficient HRP
mimic. Chemical Communications, 48, 5289–5291.
8. Roy, P., Lin, Z. H., Liang, C. T., & Chang, H. T. (2012). Synthesis of enzyme mimics of
iron telluride nanorods for the detection of glucose. Chemical Communications, 48, 4079–
4081.
9. Tian, J. Q., Liu, S., Luo, Y. L., & Sun, X. P. (2012). Fe(III)-based coordination polymer
nanoparticles: Peroxidase-like catalytic activity and their application to hydrogen peroxide
and glucose detection. Catalysis Science & Technology, 2, 432–436.
10. Liu, Q. Y., Li, H., Zhao, Q. R., Zhu, R. R., Yang, Y. T., Jia, Q. Y., et al. (2014).
Glucose-sensitive colorimetric sensor based on peroxidase mimics activity of
porphyrin-Fe3O4 nanocomposites. Materials Science & Engineering C-Materials for
Biological Applications, 41, 142–151.
11. Liu, Y., Yuan, M., Qiao, L. J., & Guo, R. (2014). An efficient colorimetric biosensor for
glucose based on peroxidase-like protein-Fe3O4 and glucose oxidase nanocomposites.
Biosensors & Bioelectronics, 52, 391–396.
12. Chang, Q., & Tang, H. Q. (2014). Optical determination of glucose and hydrogen peroxide
using a nanocomposite prepared from glucose oxidase and magnetite nanoparticles
immobilized on graphene oxide. Microchimica Acta, 181, 527–534.
13. Zhang, T. B., Lu, Y. C., & Luo, G. S. (2014). Synthesis of hierarchical iron hydrogen
phosphate crystal as a robust peroxidase mimic for stable H2O2 detection. ACS Applied
Materials & Interfaces, 6, 14433–14438.
122 Appendix

14. Ai, L. H., Li, L. L., Zhang, C. H., Fu, J., & Jiang, J. (2013). MIL-53(Fe): A metal-organic
framework with intrinsic peroxidase-like catalytic activity for colorimetric biosensing.
Chemistry-A European Journal, 19, 15105–15108.
15. Chen, W., Chen, J., Feng, Y. B., Hong, L., Chen, Q. Y., Wu, L. F., et al. (2012).
Peroxidase-like activity of water-soluble cupric oxide nanoparticles and its analytical
application for detection of hydrogen peroxide and glucose. Analyst, 137, 1706–1712.
16. Jv, Y., Li, B. X., & Cao, R. (2010). Positively-charged gold nanoparticles as peroxidiase
mimic and their application in hydrogen peroxide and glucose detection. Chemical
Communications, 46, 8017–8019.
17. Wang, X. X., Wu, Q., Shan, Z., & Huang, Q. M. (2011). BSA-stabilized Au clusters as
peroxidase mimetics for use in xanthine detection. Biosensors & Bioelectronics, 26, 3614–
3619.
18. Liu, J. B., Hu, X. N., Hou, S., Wen, T., Liu, W. Q., Zhu, X., et al. (2012). Au@Pt core/shell
nanorods with peroxidase- and ascorbate oxidase-like activities for improved detection of
glucose. Sensors and Actuators B-Chemical, 166, 708–714.
19. Wan, L. J., Liu, J. H., & Huang, X. J. (2014). Novel magnetic nickel telluride nanowires
decorated with thorns: Synthesis and their intrinsic peroxidase-like activity for detection of
glucose. Chemical Communications, 50, 13589–13591.
20. Song, Y. J., Qu, K. G., Zhao, C., Ren, J. S., & Qu, X. G. (2010). Graphene oxide: Intrinsic
peroxidase catalytic activity and its application to glucose detection. Advanced Materials,
22, 2206–2210.
21. Guo, Y. J., Li, J., & Dong, S. J. (2011). Hemin functionalized graphene nanosheets-based
dual biosensor platforms for hydrogen peroxide and glucose. Sensors and Actuators
B-Chemical, 160, 295–300.
22. Shi, W. B., Wang, Q. L., Long, Y. J., Cheng, Z. L., Chen, S. H., Zheng, H. Z., et al. (2011).
Carbon nanodots as peroxidase mimetics and their applications to glucose detection.
Chemical Communications, 47, 6695–6697.
23. Liu, S., Tian, J. Q., Wang, L., Luo, Y. L., & Sun, X. P. (2012). A general strategy for the
production of photoluminescent carbon nitride dots from organic amines and their
application as novel peroxidase-like catalysts for colorimetric detection of H2O2 and
glucose. RSC Advances, 2, 411–413.
24. Li, N., Yan, Y., Xia, B. Y., Wang, J. Y., & Wang, X. (2014). Novel tungsten carbide
nanorods: An intrinsic peroxidase mimetic with high activity and stability in aqueous and
organic solvents. Biosensors & Bioelectronics, 54, 521–527.
25. Zhang, Y. W., Tian, J. Q., Liu, S., Wang, L., Qin, X. Y., Lu, W. B., et al. (2012). Novel
application of CoFe layered double hydroxide nanoplates for colorimetric detection of
H2O2 and glucose. Analyst, 137, 1325–1328.
26. Yang, W. S., Hao, J. H., Zhang, Z., Lu, B. P., Zhang, B. L., & Tang, J. L. (2014).
CoxFe3-xO4 hierarchical nanocubes as peroxidase mimetics and their applications in H2O2
and glucose detection. RSC Advances, 4, 35500–35504.
27. Liu, Q. Y., Zhu, R. R., Du, H., Li, H., Yang, Y. T., Jia, Q. Y., et al. (2014). Higher catalytic
activity of porphyrin functionalized Co3O4 nanostructures for visual and colorimetric
detection of H2O2 and glucose. Materials Science & Engineering C-Materials for
Biological Applications, 43, 321–329.
28. Liu, S., Wang, L., Zhai, J. F., Luo, Y. L., & Sun, X. P. (2011). Carboxyl functionalized
mesoporous polymer: A novel peroxidase-like catalyst for H2O2 detection. Analytical
Methods, 3, 1475–1477.
29. Chen, X. M., Su, B. Y., Cai, Z. X., Chen, X., & Oyama, M. (2014). PtPd nanodendrites
supported on graphene nanosheets: A peroxidase-like catalyst for colorimetric detection of
H2O2. Sensors and Actuators B-Chemical, 201, 286–292.
30. Chen, X., Zhou, X. D., & Hu, J. M. (2012). Pt-DNA complexes as peroxidase mimetics and
their applications in colorimetric detection of H2O2 and glucose. Analytical Methods, 4,
2183–2187.
Appendix 123

31. Qiao, F. M., Chen, L. J., Li, X. N., Li, L. F., & Ai, S. Y. (2014). Peroxidase-like activity of
manganese selenide nanoparticles and its analytical application for visual detection of
hydrogen peroxide and glucose. Sensors and Actuators B-Chemical, 193, 255–262.
32. Zhang, W. M., Ma, D., & Du, J. X. (2014). Prussian blue nanoparticles as peroxidase
mimetics for sensitive colorimetric detection of hydrogen peroxide and glucose. Talanta,
120, 362–367.
33. Wang, T., Fu, Y. C., Chai, L. Y., Chao, L., Bu, L. J., Meng, Y., et al. (2014). Filling carbon
nanotubes with prussian blue nanoparticles of high peroxidase- like catalytic activity for
colorimetric chemo and biosensing. Chemistry-A European Journal, 20, 2623–2630.
34. Tao, Y., Ju, E. G., Ren, J. S., & Qu, X. G. (2014). Polypyrrole nanoparticles as promising
enzyme mimics for sensitive hydrogen peroxide detection. Chemical Communications, 50,
3030–3032.
35. Liu, M., Zhao, H. M., Chen, S., Yu, H. T., & Quan, X. (2012). Stimuli-responsive
peroxidase mimicking at a smart graphene interface. Chemical Communications, 48, 7055–
7057.
36. Wang, J. J., Han, D. X., Wang, X. H., Qi, B., & Zhao, M. S. (2012). Polyoxometalates as
peroxidase mimetics and their applications in H2O2 and glucose detection. Biosensors &
Bioelectronics, 36, 18–21.
37. Jiang, Z. L., Kun, L., Ouyang, H. X., Liang, A. H., & Jiang, H. S. (2011). A simple and
sensitive fluorescence quenching method for the determination of H2O2 using rhodamine B
and Fe3O4 nanocatalyst. Journal of Fluorescence, 21, 2015–2020.
38. Luo, W., Li, Y. S., Yuan, J., Zhu, L. H., Liu, Z. D., Tang, H. Q., et al. (2010). Ultrasensitive
fluorometric determination of hydrogen peroxide and glucose by using multiferroic BiFeO3
nanoparticles as a catalyst. Talanta, 81, 901–907.
39. Gao, Y., Wang, G. N., Huang, H., Hu, J. J., Shah, S. M., & Su, X. G. (2011). Fluorometric
method for the determination of hydrogen peroxide and glucose with Fe3O4 as catalyst.
Talanta, 85, 1075–1080.
40. Shi, Y., Su, P., Wang, Y. Y., & Yang, Y. (2014). Fe3O4 peroxidase mimetics as a general
strategy for the fluorescent detection of H2O2-involved systems. Talanta, 130, 259–264.
41. Hu, A. L., Liu, Y. H., Deng, H. H., Hong, G. L., Liu, A. L., Lin, X. H., et al. (2014).
Fluorescent hydrogen peroxide sensor based on cupric oxide nanoparticles and its
application for glucose and L-lactate detection. Biosensors & Bioelectronics, 61, 374–378.
42. Vdovenko, M. M., Demiyanova, A. S., Kopylov, K. E., & Sakharov, I. Y. (2014).
Fe-III-TAML activator: A potent peroxidase mimic for chemiluminescent determination of
hydrogen peroxide. Talanta, 125, 361–365.
43. He, S. H., Shi, W. B., Zhang, X. D., Li, J. A., & Huang, Y. M. (2010).
β-cyclodextrins-based inclusion complexes of CoFe2O4 magnetic nanoparticles as catalyst
for the luminol chemiluminescence system and their applications in hydrogen peroxide
detection. Talanta, 82, 377–383.
44. Shi, W. B., Zhang, X. D., He, S. H., & Huang, Y. M. (2011). CoFe2O4 magnetic
nanoparticles as a peroxidase mimic mediated chemiluminescence for hydrogen peroxide
and glucose. Chemical Communications, 47, 10785–10787.
45. Fan, Y. W., & Huang, Y. M. (2012). The effective peroxidase-like activity of
chitosan-functionalized CoFe2O4 nanoparticles for chemiluminescence sensing of hydrogen
peroxide and glucose. Analyst, 137, 1225–1231.
46. Zhang, L. H., Zhai, Y. M., Gao, N., Wen, D., & Dong, S. J. (2008). Sensing H2O2 with
layer-by-layer assembled Fe3O4-PDDA nanocomposite film. Electrochemistry
Communications, 10, 1524–1526.
47. Liu, Z. L., Zhao, B., Shi, Y., Guo, C. L., Yang, H. B., & Li, Z. A. (2010). Novel
nonenzymatic hydrogen peroxide sensor based on iron oxide-silver hybrid submicro-
spheres. Talanta, 81, 1650–1654.
124 Appendix

48. Kang, S. Z., Chen, H., & Mu, J. (2011). Electrodes modified with multiwalled carbon
nanotubes carrying Fe3O4 beads: High sensitivity to H2O2. Solid State Sciences, 13, 142–
145.
49. Kim, M. I., Ye, Y., Won, B. Y., Shin, S., Lee, J., & Park, H. G. (2011). A highly efficient
electrochemical biosensing platform by employing conductive nanocomposite entrapping
magnetic nanoparticles and oxidase in mesoporous carbon foam. Advanced Functional
Materials, 21, 2868–2875.
50. Liu, X. X., Zhu, H., & Yang, X. R. (2011). An amperometric hydrogen peroxide chemical
sensor based on graphene-Fe3O4 multilayer films modified ITO electrode. Talanta, 87,
243–248.
51. Yang, J., Xiang, H., Shuai, L., & Gunasekaran, S. (2011). A sensitive enzymeless
hydrogen-peroxide sensor based on epitaxially-grown Fe3O4 thin film. Analytica Chimica
Acta, 708, 44–51.
52. Zhang, Z. X., Zhu, H., Wang, X. L., & Yang, X. R. (2011). Sensitive electrochemical
sensor for hydrogen peroxide using Fe3O4 magnetic nanoparticles as a mimic for
peroxidase. Microchimica Acta, 174, 183–189.
53. Ye, Y. P., Kong, T., Yu, X. F., Wu, Y. K., Zhang, K., & Wang, X. P. (2012). Enhanced
nonenzymatic hydrogen peroxide sensing with reduced graphene oxide/ferroferric oxide
nanocomposites. Talanta, 89, 417–421.
54. Dutta, A. K., Maji, S. K., Srivastava, D. N., Mondal, A., Biswas, P., Paul, P., et al. (2012).
Peroxidase-like activity and amperometric sensing of hydrogen peroxide by Fe2O3 and
prussian blue-modified Fe2O3 nanoparticles. Journal of Molecular Catalysis A-Chemical,
360, 71–77.
55. Miao, Y. Q., Wang, H., Shao, Y. Y., Tang, Z. W., Wang, J., & Lin, Y. H. (2009).
Layer-by-layer assembled hybrid film of carbon nanotubes/iron oxide nanocrystals for
reagentless electrochemical detection of H2O2. Sensors and Actuators B-Chemical, 138,
182–188.
56. Fang, H. T., Pan, Y. L., Shan, W. Q., Guo, M. L., Nie, Z., Huang, Y., et al. (2014).
Enhanced nonenzymatic sensing of hydrogen peroxide released from living cells based on
Fe3O4/self-reduced graphene nanocomposites. Analytical Methods, 6, 6073–6081.
57. Dai, Z. H., Liu, S. H., Bao, J. C., & Ju, H. X. (2009). Nanostructured FeS as a mimic
peroxidase for biocatalysis and biosensing. Chemistry-A European Journal, 15, 4321–
4326.
58. Dutta, A. K., Maji, S. K., Srivastava, D. N., Mondal, A., Biswas, P., Paul, P., et al. (2012).
Synthesis of FeS and FeSe nanoparticles from a single source precursor: A study of their
photocatalytic activity, peroxidase-like behavior, and electrochemical sensing of H2O2.
ACS Applied Materials & Interfaces, 4, 1919–1927.
59. Maji, S. K., Dutta, A. K., Biswas, P., Srivastava, D. N., Paul, P., Mondal, A., et al. (2012).
Synthesis and characterization of FeS nanoparticles obtained from a dithiocarboxylate
precursor complex and their photocatalytic, electrocatalytic and biomimic peroxidase
behavior. Applied Catalysis A-General, 419, 170–177.
60. Mu, J. S., Wang, Y., Zhao, M., & Zhang, L. (2012). Intrinsic peroxidase-like activity and
catalase-like activity of Co3O4 nanoparticles. Chemical Communications, 48, 2540–2542.
61. Zhang, F. T., Long, X., Zhang, D. W., Sun, Y. L., Zhou, Y. L., Ma, Y. R., et al. (2014).
Layered double hydroxide-hemin nanocomposite as mimetic peroxidase and its application
in sensing. Sensors and Actuators B-Chemical, 192, 150–156.
62. Cui, R. J., Han, Z. D., & Zhu, J. J. (2011). Helical carbon nanotubes: Intrinsic peroxidase
catalytic activity and its application for biocatalysis and biosensing. Chemistry-A European
Journal, 17, 9377–9384.
63. Wang, Y. L., Chen, S. H., Ni, F., Gao, F., & Li, M. G. (2009). Peroxidase-like layered
double hydroxide nanoflakes for electrocatalytic reduction of H2O2. Electroanalysis, 21,
2125–2132.
Appendix 125

64. Cui, L., Yin, H. S., Dong, J., Fan, H., Liu, T., Ju, P., et al. (2011). A mimic peroxidase
biosensor based on calcined layered double hydroxide for detection of H2O2. Biosensors &
Bioelectronics, 26, 3278–3283.
65. Maji, S. K., Dutta, A. K., Srivastava, D. N., Paul, P., Mondal, A., & Adhikary, B. (2012).
Peroxidase-like behavior, amperometric biosensing of hydrogen peroxide and photocat-
alytic activity by cadmium sulfide nanoparticles. Journal of Molecular Catalysis
A-Chemical, 358, 1–9.
66. Yu, C. J., Lin, C. Y., Liu, C. H., Cheng, T. L., & Tseng, W. L. (2010). Synthesis of poly
(diallyldimethylammonium chloride)-coated Fe3O4 nanoparticles for colorimetric sensing
of glucose and selective extraction of thiol. Biosensors & Bioelectronics, 26, 913–917.
67. Mitra, K., Ghosh, A. B., Sarkar, A., Saha, N., & Dutta, A. K. (2014). Colorimetric
estimation of human glucose level using gamma-Fe2O3 nanoparticles: An easily
recoverable effective mimic peroxidase. Biochemical and Biophysical Research
Communications, 451, 30–35.
68. Lin, T. R., Zhong, L. S., Wang, J., Guo, L. Q., Wu, H. Y., Guo, Q. Q., et al. (2014).
Graphite-like carbon nitrides as peroxidase mimetics and their applications to glucose
detection. Biosensors & Bioelectronics, 59, 89–93.
69. Yu, F. Q., Huang, Y. Z., Cole, A. J., & Yang, V. C. (2009). The artificial peroxidase
activity of magnetic iron oxide nanoparticles and its application to glucose detection.
Biomaterials, 30, 4716–4722.
70. Liu, Y. P., & Yu, F. Q. (2011). Substrate-specific modifications on magnetic iron oxide
nanoparticles as an artificial peroxidase for improving sensitivity in glucose detection.
Nanotechnology, 22, 145704.
71. Su, L., Feng, J., Zhou, X. M., Ren, C. L., Li, H. H., & Chen, X. G. (2012). Colorimetric
detection of urine glucose based ZnFe2O4 magnetic nanoparticles. Analytical Chemistry,
84, 5753–5758.
72. Xing, Z. C., Tian, J. Q., Asiri, A. M., Qusti, A. H., Al-Youbi, A. O., & Sun, X. P. (2014).
Two-dimensional hybrid mesoporous Fe2O3-graphene nanostructures: A highly active and
reusable peroxidase mimetic toward rapid, highly sensitive optical detection of glucose.
Biosensors & Bioelectronics, 52, 452–457.
73. Qu, K. G., Shi, P., Ren, J. S., & Qu, X. G. (2014). Nanocomposite incorporating V2O5
nanowires and gold nanoparticles for mimicking an enzyme cascade reaction and its
application in the detection of biomolecules. Chemistry-A European Journal, 20, 7501–
7506.
74. Wang, G. L., Xu, X. F., Wu, X. M., Cao, G. X., Dong, Y. M., & Li, Z. J. (2014).
Visible-light-stimulated enzymelike activity of graphene oxide and its application for facile
glucose sensing. Journal of Physical Chemistry C, 118, 28109–28117.
75. Lin, T. R., Zhong, L. S., Guo, L. Q., Fu, F. F., & Chen, G. N. (2014). Seeing diabetes:
Visual detection of glucose based on the intrinsic peroxidase-like activity of MoS2
nanosheets. Nanoscale, 6, 11856–11862.
76. Lin, T. R., Zhong, L. S., Song, Z. P., Guo, L. Q., Wu, H. Y., Guo, Q. Q., et al. (2014).
Visual detection of blood glucose based on peroxidase-like activity of WS2 nanosheets.
Biosensors & Bioelectronics, 62, 302–307.
77. Liu, C. H., & Tseng, W. L. (2011). Oxidase-functionalized Fe3O4 nanoparticles for
fluorescence sensing of specific substrate. Analytica Chimica Acta, 703, 87–93.
78. Yang, L. Q., Ren, X. L., Tang, F. Q., & Zhang, L. (2009). A practical glucose biosensor
based on Fe3O4 nanoparticles and chitosan/nafion composite film. Biosensors &
Bioelectronics, 25, 889–895.
79. Yang, Z. P., Zhang, C. J., Zhang, J. X., & Bai, W. B. (2014). Potentiometric glucose
biosensor based on core-shell Fe3O4-enzyme-polypyrrole nanoparticles. Biosensors &
Bioelectronics, 51, 268–273.
126 Appendix

80. Niu, X. Y., Xu, Y. Y., Dong, Y. L., Qi, L. Y., Qi, S. D., Chen, H. L., et al. (2014). Visual
and quantitative determination of dopamine based on CoxFe3-xO4 magnetic nanoparticles as
peroxidase mimetics. Journal of Alloys and Compounds, 587, 74–81.
81. Zheng, C., Zheng, A. X., Liu, B., Zhang, X. L., He, Y., Li, J., et al. (2014). One-pot
synthesized DNA-templated Ag/Pt bimetallic nanoclusters as peroxidase mimics for
colorimetric detection of thrombin. Chemical Communications, 50, 13103–13106.
82. Jiang, Z. W., Liu, Y., Hu, X. O., & Li, Y. F. (2014). Colorimetric determination of thiol
compounds in serum based on Fe-MIL-88NH(2) metal-organic framework as peroxidase
mimetics. Analytical Methods, 6, 5647–5651.
83. Zhang, Z. X., Wang, X. L., & Yang, X. R. (2011). A sensitive choline biosensor using
Fe3O4 magnetic nanoparticles as peroxidase mimics. Analyst, 136, 4960–4965.
84. He, S. B., Wu, G. W., Deng, H. H., Liu, A. L., Lin, X. H., Xia, X. H., et al. (2014). Choline
and acetylcholine detection based on peroxidase-like activity and protein antifouling
property of platinum nanoparticles in bovine serum albumin scaffold. Biosensors &
Bioelectronics, 62, 331–336.
85. Qian, J., Yang, X. W., Jiang, L., Zhu, C. D., Mao, H. P., & Wang, K. (2014). Facile
preparation of Fe3O4 nanospheres/reduced graphene oxide nanocomposites with high
peroxidase-like activity for sensitive and selective colorimetric detection of acetylcholine.
Sensors and Actuators B-Chemical, 201, 160–166.
86. Shamsipur, M., Safavi, A., & Mohammadpour, Z. (2014). Indirect colorimetric detection of
glutathione based on its radical restoration ability using carbon nanodots as nanozymes.
Sensors and Actuators B-Chemical, 199, 463–469.
87. Kim, M. I., Shim, J., Li, T., Woo, M. A., Cho, D., Lee, J., et al. (2012). Colorimetric
quantification of galactose using a nanostructured multi-catalyst system entrapping
galactose oxidase and magnetic nanoparticles as peroxidase mimetics. Analyst, 137,
1137–1143.
88. Ni, P. J., Dai, H. C., Wang, Y. L., Sun, Y. J., Shi, Y., Hu, J. T., et al. (2014). Visual
detection of melamine based on the peroxidase-like activity enhancement of bare gold
nanoparticles. Biosensors & Bioelectronics, 60, 286–291.
89. Sharma, T. K., Ramanathan, R., Weerathunge, P., Mohammadtaheri, M., Daima, H. K.,
Shukla, R., et al. (2014). Aptamer-mediated ‘turn-off/turn-on’ nanozyme activity of gold
nanoparticles for kanamycin detection. Chemical Communications, 50, 15856–15859.
90. Sun, Z. Z., Zhang, N., Si, Y. M., Li, S., Wen, J. W., Zhu, X. B., et al. (2014).
High-throughput colorimetric assays for mercury(II) in blood and wastewater based on the
mercury-stimulated catalytic activity of small silver nanoparticles in a
temperature-switchable gelatin matrix. Chemical Communications, 50, 9196–9199.
91. Mohammadpour, Z., Safavi, A., & Shamsipur, M. (2014). A new label free colorimetric
chemosensor for detection of mercury ion with tunable dynamic range using carbon
nanodots as enzyme mimics. Chemical Engineering Journal, 255, 1–7.
92. Wu, G. W., He, S. B., Peng, H. P., Deng, H. H., Liu, A. L., Lin, X. H., et al. (2014).
Citrate-capped platinum nanoparticle as a smart probe for ultrasensitive mercury sensing.
Analytical Chemistry, 86, 10955–10960.
93. Mu, J. S., Zhang, L., Zhao, M., & Wang, Y. (2014). Catalase mimic property of Co3O4
nanomaterials with different morphology and its application as a calcium sensor. ACS
Applied Materials & Interfaces, 6, 7090–7098.
94. Gao, L. Z., Zhuang, J., Nie, L., Zhang, J. B., Zhang, Y., Gu, N., et al. (2007). Intrinsic
peroxidase-like activity of ferromagnetic nanoparticles. Nature Nanotechnology, 2, 577–
583.
95. Gao, L. Z., Wu, J. M., Lyle, S., Zehr, K., Cao, L. L., & Gao, D. (2008). Magnetite
nanoparticle-linked immunosorbent assay. Journal of Physical Chemistry C, 112, 17357–
17361.
Appendix 127

96. Zhang, X. Q., Gong, S. W., Zhang, Y., Yang, T., Wang, C. Y., & Gu, N. (2010). Prussian
blue modified iron oxide magnetic nanoparticles and their high peroxidase-like activity.
Journal of Materials Chemistry, 20, 5110–5116.
97. Tang, Z. W., Wu, H., Zhang, Y. Y., Li, Z. H., & Lin, Y. H. (2011). Enzyme-mimic activity
of ferric nano-core residing in ferritin and its biosensing applications. Analytical Chemistry,
83, 8611–8616.
98. Bhattacharya, D., Baksi, A., Banerjee, I., Ananthakrishnan, R., Maiti, T. K., & Pramanik,
P. (2011). Development of phosphonate modified Fe(1-x)MnxFe2O4 mixed ferrite nanopar-
ticles: Novel peroxidase mimetics in enzyme linked immunosorbent assay. Talanta, 86,
337–348.
99. Figueroa-Espi, V., Alvarez-Paneque, A., Torrens, M., Otero-Gonzalez, A. J., & Reguera, E.
(2011). Conjugation of manganese ferrite nanoparticles to an anti Sticholysin monoclonal
antibody and conjugate applications. Colloids and Surfaces A-Physicochemical and
Engineering Aspects, 387, 118–124.
100. Kumari, S., Dhar, B. B., Panda, C., & Meena, A. (2014). Sen Gupta, S. Fe-TAML
encapsulated inside mesoporous silica nanoparticles as peroxidase mimic: Femtomolar
protein detection. ACS Applied Materials & Interfaces, 6, 13866–13873.
101. Dong, J. L., Song, L. N., Yin, J. J., He, W. W., Wu, Y. H., Gu, N., et al. (2014). Co3O4
nanoparticles with multi-enzyme activities and their application in immunohistochemical
assay. ACS Applied Materials & Interfaces, 6, 1959–1970.
102. Song, Y. J., Xia, X. F., Wu, X. F., Wang, P., & Qin, L. D. (2014). Integration of platinum
nanoparticles with a volumetric bar-chart chip for biomarker assays. Angewandte
Chemie-International Edition, 53, 12451–12455.
103. Zhang, L. N., Deng, H. H., Lin, F. L., Xu, X. W., Weng, S. H., Liu, A. L., et al. (2014). In
situ growth of porous platinum nanoparticles on graphene oxide for colorimetric detection
of cancer cells. Analytical Chemistry, 86, 2711–2718.
104. Zhan, L., Li, C. M., Wu, W. B., & Huang, C. Z. (2014). A colorimetric immunoassay for
respiratory syncytial virus detection based on gold nanoparticles-graphene oxide hybrids
with mercury-enhanced peroxidase-like activity. Chemical Communications, 50, 11526–
11528.
105. Hu, X. N., Saran, A., Hou, S., Wen, T., Ji, Y. L., Liu, W. Q., et al. (2014). Rod-shaped
Au@PtCu nanostructures with enhanced peroxidase-like activity and their ELISA
application. Chinese Science Bulletin, 59, 2588–2596.
106. He, W. W., Liu, Y., Yuan, J. S., Yin, J. J., Wu, X. C., Hu, X. N., et al. (2011). Au@Pt
nanostructures as oxidase and peroxidase mimetics for use in immunoassays. Biomaterials,
32, 1139–1147.
107. Qu, F. L., Li, T., & Yang, M. H. (2011). Colorimetric platform for visual detection of
cancer biomarker based on intrinsic peroxidase activity of graphene oxide. Biosensors &
Bioelectronics, 26, 3927–3931.

You might also like