PreprintPDF Available

Biosensors for the detection of bacterial and viral clinical pathogens and covid-19 diagnosis v2.0

Authors:

Abstract

Biosensors are measurement devices that can sense several biomolecules, and are widely used for the detection of relevant clinical pathogens such as bacteria and viruses, showing outstanding results. Because of the latent existing risk of facing another pandemic like the one we are living due to COVID-19, researchers are constantly looking forward to developing new technologies for diagnosis and treatment of infections caused by different bacteria and viruses. Regarding that, nanotechnology has improved biosensors design and performance through the development of materials and nanoparticles that enhance their affinity, selectivity, and efficacy in detecting these pathogens, such as employing nanoparticles, graphene quantum dots, and electrospun nanofibers. Therefore, this work aims to present a comprehensive review that exposes how biosensors work in terms of bacterial and viral detection, and the nanotechnological features that are contributing to achieving a faster yet still efficient COVID-19 diagnosis at the point-of-care.
Review
Biosensors for the Detection of Bacterial and Viral
Clinical Pathogens and COVID-19 Diagnosis
Luis Castillo-Henríquez1,2, Mariana Brenes-Acuña3, Arianna
Castro-Rojas3, Rolando Cordero-Salmerón3, Mary Lopretti4 and
José Roberto Vega-Baudrit 1,3,*
1National Laboratory of Nanotechnology (LANOTEC), National Center for High Technology
(CeNAT), 1174-1200, San José, Costa Rica; luis.castillohenriquez@ucr.ac.cr (L.C.H)
2Physical Chemistry Laboratory, Faculty of Pharmacy, University of Costa Rica, 11501-2060, San
José, Costa Rica
3Chemistry School, National University of Costa Rica, 86-3000, Heredia, Costa Rica;
mbrenesacua@gmail.com (M.B.A); ariannac.r9vi@gmail.com (A.C.R); rocordero105@gmail.com
(R.C.S)
4UDELAR University, 1140, Montevideo, Uruguay; mlopretti@gmail.com (M.L.C)
*Correspondence: jvegab@gmail.com
Received: date; Accepted: date; Published: date
Abstract: Biosensors are measurement devices that can sense several biomolecules, and are widely
used for the detection of relevant clinical pathogens such as bacteria and viruses, showing
outstanding results. Because of the latent existing risk of facing another pandemic like the one we
are living due to COVID-19, researchers are constantly looking forward to developing new
technologies for diagnosis and treatment of infections caused by different bacteria and viruses.
Regarding that, nanotechnology has improved biosensors design and performance through the
development of materials and nanoparticles that enhance their affinity, selectivity, and efficacy in
detecting these pathogens, such as employing nanoparticles, graphene quantum dots, and
electrospun nanofibers. Therefore, this work aims to present a comprehensive review that exposes
how biosensors work in terms of bacterial and viral detection, and the nanotechnological features
that are contributing to achieving a faster yet still efficient COVID-19 diagnosis at the point-of-care.
Keywords: Bacterial detection; Biosensors; Clinical pathogen; COVID-
19; Electrospun nanofibers; Nano-biosensors; Point-of-care; SARS-CoV-2;
Viral detection.
1. Introduction
Biosensor’s concept was firstly addressed by Clark and Lyons around 1962 when they
developed an oxidase enzyme electrode for glucose detection [1]. Since then, nanotechnological
development has promoted biosensors evolution and specialization for different purposes [2].
Sensors 2020, 20, x; doi: FOR PEER REVIEW www.mdpi.com/journal/sensors
1
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
2
Sensors 2020, 20, x FOR PEER REVIEW 2 of 32
Currently, nanotechnology is at the forefront of science, and its combination with biosensoring
applications involves different fields such as medicine, biology, environmental, drug delivery, food
safety, and others [3–7]. However, the detection of pathogens has become one of the most relevant
objectives for these devices since bacterial and viral diseases currently represent an important
thread for human health [8,9].
Virus and bacteria detection commonly involves the use of several molecular techniques
such as the reverse transcription-polymerase chain reaction (RT-PCR), which remains the gold
standard for pathogen detection [10]. The classical detection methods for these pathogens usually
require isolation, culturing, and then, biochemical tests [11]. Additionally, serological tests like
ELISA are used for the detection of antibodies and immunoglobulin needed for identification
purposes [12]. However, some of these techniques take a long time for obtaining results and are
usually laborious. Therefore, new approaches based on nanotechnological advances have emerged
as suitable and easier options for detecting pathogens in faster and efficient ways [11,13].
On one hand, nanoparticles (NPs) have demonstrated outstanding properties against
different pathogens used to develop novel devices and technologies that contribute to this public
health issue [14,15]. The interest is not limited to human diseases, but also considers the ones
affecting animals since zoonosis is an existent thread. Stringer et al. developed an optical biosensor
using gold NPs (AuNPs) and quantum dots (QDs) for the detection of porcine reproductive and
respiratory syndrome virus [16].
On the other hand, international scientific community’s interest in using DNA biosensors
or sequence-specific DNA detectors for clinical studies is increasingly growing. In 2007, Dell’Atti et
al. developed a combined DNA-based piezoelectric biosensor for simultaneous detection and
genotyping of high-risk Human Papilloma Virus (HPV) strains [17]. In addition to that, these
biosensors have been employed for DNA damage research and specific gene sequences detection
[18,19].
Biosensors and nano-biosensors have been extensively used for the detection of viral and
bacterial clinical pathogens. These devices are practical (e.g., enable point-of-care (POC) testing
through smartphone-based nano-biosensor), fast, and are considered as innovative technologies
that provide an alternative solution to the mentioned disadvantages presented by common
detection methods [20–22]. The aforementioned have been employed for studying viruses affecting
human health such as Ebola virus, Human Immunodeficiency Virus (HIV), and more recently the
newly discovered acute respiratory syndrome coronavirus 2 (SARS-CoV-2), as well as bacteria like
Escherichia coli, Salmonella spp. and others [23–27].
Because of the latent existing risk of facing another pandemic like we are living due to
Coronavirus disease (COVID-19), researchers are constantly looking forward to developing new
technologies for diagnosis and treatment of infections caused by different bacteria and viruses.
Therefore, this review aims to expose how biosensors work in terms of bacterial and viral detection,
describing the nanotechnological features such as NPs, graphene QDs (GQDs), and electrospun
nanofibers, which enhance their affinity, selectivity, and efficacy in detecting these pathogens, as
well as highlighting current advances for the COVID-19 pandemic assessment at the POC.
2. Biosensors
3
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
4
Sensors 2020, 20, x FOR PEER REVIEW 3 of 32
Biosensors can be defined as a measurement system for analyte detection that combines a
biological component with a physicochemical detector [28]. The analyte detection depends on the
biosensor design and purpose. Some commonly used devices such as smartphones can be
employed as a biosensor with the inclusion of simple accessories such as published by Soni et al.,
where they developed a non-invasive smartphone-based biosensor for urea using saliva as sample
[29,30]. This allows fast and low-cost preliminary detection [31].
Usually, biosensors detect biomolecules such as nucleic acids, proteins, and cells that are
associated with diseases. This is possible because of their three major components: the biologically
sensitive element, the detector element, and the reader device [32]. Enzymes, microorganisms,
organelles, antibodies, and nucleic acids are used to detect the biomolecules [33]. In order to obtain
a high-quality biosensor, researchers must identify the requirements to obtain a fully functional
device. Hence, multidisciplinary studies are fundamental to select the proper material, transducing
device, and biological element involved, before assembling the biosensor [34].
At a clinical level, biosensors are applied for detecting disease-associated biomolecules [32].
These devices can monitor several parameters like disease-causing bacteria, and several body fluids
such as saliva, blood, or urine [35,36]. Zhang et al. developed a non-invasive method for glucose
testing based on a disposable saliva nano-biosensor to improve patient compliance, reduce
complications, and costs derived from diabetes management. In the clinical trials, they obtained
outstanding results in terms of accuracy compared to the UV spectrophotometer. Thus, the
disposable device can be presented as an alternative for real-time salivary glucose tracking [37].
Biosensors can be applied for many other clinical diagnostic purposes, such as cholesterol,
markers related to cardiovascular diseases, biomarkers of cancer or tumors, allergic responses,
disease-causing bacteria, viruses, and fungi infections [38–41]. Aside from that, biosensors can be
employed for bacteria and virus detection in food and water, which are potential sources of
diseases [42,43]. Zhao et al. fabricated a low-cost, portable microfluidic chemiresistive biosensor
based on monolayer graphene, AuNPs, and streptavidin-antibody system for the rapid in-situ
detection of E. coli. In this case, the bacteria are captured on the biosensor’s surface and detection is
performed through electric readouts [44]. Another approach published by Samanman et al.
describes the development of a glutathione-S-transferase tag for white spot binding protein (GST-
WBP) immobilized onto a gold electrode through a self-assembled monolayer. This biosensor can
detect white spot syndrome virus (WSSV) in shrimp pond water due to binding between WSSV and
the immobilized GST-WBP [45].
2.1. Operating principles
Biosensors are constituted by three components (Figure 1) [38,46]. In the first place, these
devices have sensing elements, also called bioreceptor that emulates in vivo molecular recognition
phenomena [47]. There is a wide range of sensing elements such as cells, microbes, cell receptors,
antibodies, enzymes, or nucleic acids [48–52]. These biological sensitive elements recognize the
analyte and interact with it in different ways according to the type of biosensor [53]. One of the
main biorecognition strategies is based on bacterial or viral nucleic acid sequences [54,55]. Solanki
et al. developed a DNA bioelectrode to detect Vibrio cholerae, which is stable for at least 15 weeks
under 4°C storage. The biosensor consisted of O1 gene-based 24mer single-stranded DNA probe
immobilized onto solgel derived nanostructured zirconium oxide (NanoZrO2) film [56].
5
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
6
Sensors 2020, 20, x FOR PEER REVIEW 4 of 32
Figure 1. Biosensor’s basic design. Reprinted with permission from Huang, Y. et al. Disease-Related Detection
with Electrochemical Biosensors: A Review. Sensors 17(10). Copyright (2017) MDPI [46].
The second element is the transductor or detector, which works by sensing a signal related
to a physicochemical change caused by the interaction between the bioreceptor and the analyte. It
transforms the signal into another one that can be evaluated and quantified [57–61]. The last part of
a biosensor is the reader device. It usually involves a display that depends on software and
hardware to generate the results [62].
Some important attributes define the performance of a biosensor. In the first place,
selectivity is the capacity of a bioreceptor to detect a specific bio-entity when analyzing a sample
composed of other components. This is probably the main feature and determines the needed
bioreceptor. Second, reproducibility is the ability to produce the same response for a certain
experimental set-up that is performed multiple times. Reproducible signals provide high reliability
and robustness. Third, stability is the capacity to endure ambient disturbances around the system
that can affect the precision and accuracy of the device. Fourth, sensitivity also known as the limit
of detection (LOD) is the minimum amount of the analyte that can be detected by a biosensor. For
clinical applications, it is required to detect the analyte in samples of low concentrations (ng/ml or
fg/ml). Finally, linearity examines how accurate are the measurements within the analyte range of
concentrations (i.e. linear range), and in response to the smallest variation in terms of concentration
that can cause a change in the output (i.e. resolution) [63].
2.2. Types of biosensors
Biosensors can be classified by the way they transduce signals into optical, electrochemical,
and piezoelectric devices [57–61,64]. Optical biosensors are those that perform their analysis
through the measure of photons, using optic fibers as transduction elements [58,59,65]. Several optic
sensing mechanisms can be employed by this type of biosensor for analyte detection such as
absorption, colorimetry, fluorescence, or luminescence [66]. This kind of biosensor presents a lower
noise and immunity to electromagnetic interference, which gives it an advantage over
electrochemical and piezoelectric biosensors [67].
Vidal et al. developed a chromatic biosensor for quick bacterial detection based on
polyvinyl butyrate-polydiacetylene non-woven fiber composites. The device shows promising
potential to alert about possible infections caused by Staphylococcus aureus, Micrococcus luteus, and
E.coli [68]. In another study, Jeong et al. constructed a fluorescent supramolecular biosensor for
bacterial detection. The binding of these pathogens induces conformational changes in the
supramolecular state, which causes a fluorescence emission that can selectively detect E.coli over
7
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
8
Sensors 2020, 20, x FOR PEER REVIEW 5 of 32
other microorganisms [69]. Regarding viral analysis, Ahmadi et al. evaluated single virus detection
through an optical biosensor, where viral particles attached to a microsphere optical resonator’s
surface caused a shift of resonance to longer wavelengths [70].
The second type, electrochemical biosensor, has been extensively applied to pathogen
detection. These devices sense the analyte through electrodes by measuring electrical signals
resulting from catalytic reactions or specific unions. The previous is derived from the capture of
electrons as a result of redox reactions between the analyte and the bio-element [71]. In addition to
that, the analysis of the desired element is determined by different readouts like potentiometry,
amperometry, and conductometry [72]. This type of biosensor has been subjected to improvements
due to bio-and nanomaterials development [72,73].
Recently, Mathelié et al. employed non-cytotoxic silica NPs-assisted electrochemical
biosensor for sensitive and specific detection of E. coli. The electrochemical immune-biosensor
detects the bacteria in five minutes by cyclic voltammetry measurements, and also represents a
potential device for targeting a variety of other microorganisms through little modifications within
its features [74]. In another study, Baek et al. developed an electrochemical biosensor composed of
eight novel peptides separately in a gold electrode for the detection of human norovirus. The
peptides exhibited a high binding affinity towards the viruses, and a decrease in current signals
explained by increasing concentration of the virus [75].
Finally, yet importantly, there are piezoelectric biosensors. Piezoelectricity refers to the
ability of a material to generate a voltage under mechanical stress [76]. These biosensors possess
crystals that vibrate under the influence of an electric field. Besides, certain materials vibrate at
characteristic resonant frequencies in response to interaction with other molecules. The relationship
between the resonant frequency changes and the mass from the molecules adsorbed or desorbed
from the crystal’s surface is conceived as the working principle of transduction in this type of
biosensor. Therefore, vibration provides information on the phenomenon that is being measured
[77,78].
Fu et al. discuss the advances in piezoelectric thin films acoustic wave devices for bacterial
and viral detection of pathogens adsorbed on surfaces through DNA interaction with
complementary strands. The previous allows early detection of clinical pathogens, and thus,
prevents the spreading of the infection [79]. In another approach, Guo et al. worked on sensitive E.
coli O157:H7 detection system using a piezoelectric biosensor-quartz crystal microbalance with
antibody-functionalized AuNPs to enhance changes in detection signals. It was demonstrated that
the developed device can be used as a suitable real-time monitoring method for the mentioned
pathogen [80].
3. Biosensors nanotechnological features for bacterial and viral detection
Over time, many techniques and methods have been developed for detecting pathogens
such as viruses and bacteria, including colorimetric methods, fluorescence polarization, and
electrochemical analysis [81]. However, those are very expensive and possess limitations related to
time-consumption, low precision of the results, poor stability, and short life span [82].
Bacterial and viral outbreaks have caused many issues in biomedical, food and
environmental context, making necessary the development of new strategies that allow faster
detection of these pathogens to effectively contain and control their impact on human health [83].
9
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
10
Sensors 2020, 20, x FOR PEER REVIEW 6 of 32
The combination of nanotechnologies and biosensors’ characteristics is currently being considered
as a potential opportunity for speeding up the development of fast, highly sensitive, and specific
devices for genuine bacterial and viral detection. As a consequence, nano-biosensors make use of
chemical, electrical, optical, and magnetic properties of materials for detecting biomolecules and
pathogens [84,85].
In order to satisfy the previous, nanotechnology has greatly contributed to the development
of biosensors due to research in nanomaterials and nanostructures, such as carbon nanotubes,
GQDs, metal oxide NPs, metal nanoclusters, plasmonic nanomaterials, polymer nanocomposites,
nanogels, among others (Figure 2) [86–89]. These have been employed for modifying electrode
surfaces to improve critical features, such as reproducibility, selectivity, and sensitivity, due to their
biocompatible character, structural compatibility, and high adsorption capacity. Therefore,
nanomaterials have demonstrated to be suitable for biosensing applications, enhancing the
performance with increased sensitivities and lower detection limits [90].
Figure 2. Different nanomaterials and nanostructures used for the development of nano-biosensors. Reprinted
with permission from Pirzada, M. et al. Nanomaterials for Healthcare Biosensing Applications. Sensors 19(23):
5311. Copyright (2019) MDPI [91].
Additionally, nanomaterials have been used to increase the immobilized bioreceptor
loadings. However, the strategy for immobilizing the bio-specific entity onto the nanomaterial is
considered as the biggest challenge for developing high quality and reliable nano-biosensor. Non-
covalent approaches such as electrostatic interactions, polymers entrapment, or van der Waals
forces between the nanomaterial and the biomolecule do not alter their specific properties. On the
other hand, covalent binding provides more stability and reproducibility of surface
functionalization, as well as reducing the risk of unspecific physisorption. Although the previous
techniques represent good strategies for binding biological species to surfaces, supramolecular
interactions have recently been considered as superior since these are reversible, which enables the
regeneration of the transducer element [91,92].
Regarding other uses, nanomaterials can perform as nanocarriers for signaling elements, as
well as signal amplification. Depending on the chemical composition, nanomaterials can be subject
to direct functionalization during synthesis, or functionalized by coating using functional polymers
11
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
12
Sensors 2020, 20, x FOR PEER REVIEW 7 of 32
[93]. Nanomaterials functionalization provides three important advantages: reproducible
immobilization of bioreceptor units, increase the biocompatibility, and the development of label-
free transduction techniques [92].
Moreover, nano-biosensor materials’ high surface area is considered a major advantage
compared to conventional devices and plays an important role in the sensitivity, and fast response
of the devices [94,95]. Therefore, these are conceived as excellent tools used for the detection,
function, and interaction of proteins and nucleic acids, which improve the quality and performance
of diagnosis for bacterial and viral diseases [96]. The following sections present an overview of
some promising nanotechnological features in biosensors.
3.1. NPs
NPs are a wide range of materials with dimensions below 100 nm that have been used in
various areas such as medical, pharmaceutical, manufacturing and materials, environmental,
electronics, and mechanical industries due to their multiple properties [97–100]. Among the mostly
employed are metal NPs such as AuNPs and silver NPs (AgNPs), which can be produced in
different sizes and shapes (e.g., nanospheres, nanocylinders, nanowires, and nanocages). These NPs
exhibit low toxicity, as well as multiple interesting chemical, biological, and physical properties,
such as photo-thermal, optical, electrochemical, and biocompatibility based on their inert nature in
biological fluids [101–103]. Additionally, these NPs can be synthesized with ease fulfilling relevant
roles for diagnostic probes, and functionalized due to the presence of functional groups for
achieving ligand-binding functions with a wide range of molecules, such as antibodies or genetic
material [104,105].
An important application of nano-biosensors composed of metal NPs is related to
waterborne diseases, where the infection is usually linked to microbial contamination due to several
pathogens, including bacteria, where nanotechnological detection systems with optical sensing
have been used for these pathogens [106]. Elahi et al. designed a highly sensitive fluorescence nano-
biosensor for the detection of Shigella species. To achieve a satisfactory design, two DNA probes as
sensing elements were immobilized on the surface of AuNPs synthesized for the development,
forming a DNA-probe AuNPs-fluorescence system. The research group also synthesized iron NPs
(MNPs) that were later modified with Sulfosuccinimidyl 4-Nmaleimidomethyl cyclohexane-1-
carboxylate (SMCC), and a second system constituted by a third DNA probe immobilized on MNPs
was formed for separating target DNA. The results exhibited an increasing fluorescence intensity
with an increase of target DNA concentration [107].
In another study, carried out by Takemura et al., an ultrasensitive, rapid, and specific
localized surface plasmon resonance (LSPR)-induced immunofluorescence nano-biosensor was
developed for detecting influenza virus. Researchers employed AuNPs-induced QD fluorescence
signal conjugated with antineuraminidase antibody (Anti-NA Ab) and conjugation of anti-
hemaglutinin antibody (anti-HA Ab) to the QDs. The device successfully detected influenza virus
H1N1. However, due to its versatility, it was also possible to detect clinically isolated influenza
virus H3N2 and norovirus-like particles [108].
3.2. GQDs
GQDs are among the most fascinating carbon-based nanomaterials employed for the
development of biosensors, mostly electrochemical. These materials present outstanding properties
13
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
14
Sensors 2020, 20, x FOR PEER REVIEW 8 of 32
such as signal amplifying characteristics, biocompatibility, tunable size, electro-catalytic
performance, and capacity to detect multiple biomolecules. Additionally, their inertness, non-
toxicity, long-term chemical stability, and water stability make them very valuable for biomedical
applications [90,93].
GQDs obtained through different synthesis methods have been used for biosensing
applications since their large surface area can be functionalized, allowing them to directly detect
DNA, enzymes, proteins, antigens, antibodies, and other biomolecules by the oxide components
formed on their surface during the synthesis process [109]. Safardoust et al. synthesized GQDs from
citric acid and ethylene diamine for their use as a photoluminescence sensor for detecting S.aureus
and E.coli. This biosensor demonstrated a linear relationship between the fluorescence intensity and
the concentrations of the bacterias up to 9x107 CFU/ml [110].
In another approach, Hazani et al. fabricated a highly sensitive electrochemical peptide
nucleic acid (PNA) biosensor based on functionalized graphene oxide composited with cadmium
sulfide QDs (CdS QDs). The device was developed for detecting Mycobacterium tuberculosis and
showed a LOD of 8.948x10-13 M [111]. Furthermore, GQDs integration into a biosensor can improve
its performance in terms of reproducibility, selectivity, and sensitivity [112].
3.3. Electrospun nanofibers
Electrospinning is a nanotechnological method in which an electrostatic field force applied
to a polymer solution causes a charged liquid jet to moves downfield towards an oppositely
charged collector, where the fine fibers are deposited [113]. Electrospun nanofibers have been the
target of different applications like drug delivery systems or scaffolds for skin tissue engineering
due to their structure and physicochemical properties such as a large surface area to volume ratio,
small particle size, and high porosity, among others [113–116]. However, a novel application is their
use for developing nano-biosensors focused on detecting viral and bacterial pathogens [117–120].
Nano-biosensors development using these nanostructures can be achieved by two
approaches. On one hand, functional polymers are electrospun to obtain a nanofiber that is used
directly as an inducing element of the corresponding biosensor, which will present fast response
time, high sensitivity, and good biocompatibility. On the other hand, electrospun nanofibers are
used as templates to which a sensitive material is deposited on their surface, and later the system is
subjected to chemical modification in order to produce a composite film on an electrode, with
nanostructures that have the intended sensing characteristics [121,122].
Although the manufacturing process is simpler, keeping bio-receptor functionality is
considered a great challenge for the production of this type of device. The sensing element can be
immobilized through different strategies according to its physicochemical characteristics, as well as
the ones from the nanofiber scaffolds, and also, based on their interfacial interactions [123].
Moreover, this type of nano-biosensor is based on various sensing principles such as optics,
electric resistance, photoelectricity, vibration frequency, electric current, and others [124–128]. Luo
et al. developed a nitrocellulose electrospun nanofibrous capture membrane for detecting
E.coli O157:H7 and bovine viral diarrhea virus. The device’s design was based on capillary
separation, and conductometric immunoassay using a silver electrode. Nanofiber antibody’s
surface functionalization and sensor assembly process allowed retaining the unique fiber
15
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
16
Sensors 2020, 20, x FOR PEER REVIEW 9 of 32
morphology, and displaying a linear response to both pathogens with a detection time of 8 minutes
[129].
Quiros et al. prepared electrospun membranes composed of polyacrylonitrile (PAN) and
poly(4-vinylphenylboronic acid-co-2-(dimethylamino)ethyl methacrylate-co-n-butyl methacrylate)
(pVDB) for fast sensing of bacteria. The pVDB@PAN membranes were used as fluorescent bacterial
biosensors, displaying maximum fluorescence intensity after 24 hours in contact with S. aureus or E.
coli. Meanwhile, the membranes became non-responsive within 8 hours in contact
with Pseudomonas putida due to the rapid formation of bacterial biofilm that blocked the membrane
surface, disrupting fluorescence readings. This development can be useful for the early
identification of pathogenic bacteria as an attempt to prevent their spreading [130].
Some research groups have designed nano-biosensors based on electrospun nanofibers for
viral detection as well. Tripathy et al. worked on an ultrasensitive electrochemical platform with
electrospun semi-conducting Manganese (III) Oxide (Mn2O3) nanofibers for DNA hybridization
detection. This biosensor makes use of electrochemical transduction techniques for zeptomolar (i.e.,
10-21 M) detection of Dengue primer, resulting in a limit of detection of 120×10–21 M [131].
Therefore, nanofiber-based biosensors present advantages over the conventional ones such
as polymer diversity for its manufacture, high specific surface area with high responsiveness, as
well as an outstanding sensibility [132–134].
4. Bacterial and viral pathogens detected through biosensors and nano-
biosensors
Conventional clinical analyses including an antibody or nucleic acid-based, biochemical,
and enzymatic methods, are very reliable but take a long time to obtain a result. Health disciplines
demand the acquisition of faster outcomes to speed up the appropriate treatment [135,136]. In this
sense, biosensors and nano-biosensors are useful tools that offer an accurate diagnosis in shorter
times due to their ability to provide real-time and faster clinical results [137]. Currently, there is an
increasing interest in their use to detect pathogens in the human body (Table 1) [136].
Table 1. Developed biosensors for detecting bacterial and viral pathogens in the human body.
Device Target pathogen LOD Response
time Ref
Long-period fiber grating using
bacteriophage T4 covalently
immobilized on optical fiber
surface.
E.coli 103 CFU/ml 20 min [138]
Label free polyaniline based
impedimetric. E.coli O157:H7 102 CFU/ml - [139]
Electrochemical biosensor using
antibody-modified NPs
(polymer-coated magnetic NPs
and carbohydrate-capped
AuNPs).
E.coli O157:H7 101 CFU/ml 45 min [140]
17
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
18
Sensors 2020, 20, x FOR PEER REVIEW 10 of 32
Surface plasmon resonance (SPR)
biosensor based on ultra-low
fouling and poly(carboxibetaine
acrylamide).
Salmonella sp. 7.4x103 CFU/ml 80 min [141]
Graphene-based potentiometric. S. aureus 1 CFU/ml 10-15 min [142]
Aptamer based biosensor and
dual florescence resonance
energy transfer from QDs to
carbon NPs.
Vibrio
parahaemolyticus
and Salmonella
typhimurium
25 CFU/ml and 35
CFU/ml,
respectively
80 min [112]
Impedimetric biosensor based on
site specifically attached
engineered antimicrobial
peptides.
Pseudomona
aeruginosa 102 CFU/ml 30 min [143]
Electrochemical DNA biosensor
based on flower-like ZnO
nanostructures.
Neisseria
meningitides 5 ng/μl - [144]
Graphene-enabled biosensor
with a highly specific
immobilized monoclonal
antibody.
Zika virus 0.45 nM 4-8 min [145]
Giant magnetoresistance
biosensor. Influenza A virus 1.5x102 TCID50/mL - [146]
Electrochemical biosensor based
on DNA hybridization. Hepatitis A virus 6.94 fg/μl 15 min [147]
Impedimetric electrochemical
DNA biosensor for label free
detection.
Zika virus 25 nM 1.5 h [148]
Two-dimensional molybdenum
disulphide nanosheets based
disposable biosensor.
Chikungunya virus 3.4 nM 3 h [149]
Electrochemical DNA biosensor
using gold nanorods. Hepatitis B virus 2.0x10-12 mol/L 5 h [150]
Intensity-modulated surface
plasmon resonance (IM-SPR)
biosensor
Avian influenza A
H7N9 virus 144 copies/ml 10 min [151]
Silicon nanowire biosensor. Dengue virus 2.0 fM - [152]
E.coli: Escherichia coli; S.aureus: Staphylococcus aureus.
Molecular determination demands to improve the analytical performance of biosensors,
which have enhanced their unique features to develop POC devices in order to run a rapid and
cost-effective analysis of complex biological matrices [153]. Commercial versions of these devices
are available to detect diseases and pathogens such as E. coli, Helicobacter pylori, influenza A and B
viruses, HIV, tuberculosis, and malaria [154]. Advantages such as small samples and low energy
required to avoid complications in terms of transportation and processing, make them suitable for
19
342
343
344
345
346
347
348
20
Sensors 2020, 20, x FOR PEER REVIEW 11 of 32
easy and fast use in the identification of bacterial and viral pathogens [137]. Needless to say,
nanomaterials advances have benefited biosensor performance to achieve the task [155].
4.1. Bacterial pathogen detection
Focusing on the human body, bacterial infections caused mainly by gram-negative
microorganisms represent a particular challenge in human health worldwide because of multidrug
resistance variants, greatly influenced by their indiscriminate exposure to antibiotics discharged in
water, addition to food or more commonly, due to improper use of these drugs from patients [156].
Since the previously mentioned is considered a major current health concern, different
kinds of nanomaterials and biorecognition elements have been employed to develop biosensors for
antibiotic detection, as well for bacteria [157]. Common pathogenic bacteria include E. coli,
Salmonella typhi, Clostridium perfringens, and Shigella spp., which can cause different kinds of diseases
in humans, animals, and plants [158]. However, S. aureus is recognized as one of the most fatal
bacteria that can cause rapid mortal infections and is often resistant to multiple antibacterial active
substances. Thus, it is necessary to develop new approaches for easier and faster detection since
conventional culture methods require 3-5 days to obtain results, and other nucleic acid-based
methods are expensive and imply trained personnel [159,160].
Suaifan et al. developed a biosensor able to detect S. aureus in a few minutes. The sensing
tool is based on the proteolytic activity of the pathogen proteases on a specific peptide substrate
placed in the middle of two magnetic nanobeads. In this case, the dissociation of magnetic
nanobeads-peptide moieties results in color change [161]. In another approach, Ahari et al.
constructed a potentiometric nano-biosensor able to detect the bacteria through the identification of
an exotoxin emitted by the microorganism. Particularly, the method is often used for contaminated
food, but it can also be applied for clinical detection [162].
Another important bacteria, V. cholerae, is a gram-negative facultative anaerobe that causes
Cholera disease. People would infect by consuming contaminated liquids or food, providing an
ideal platform for the disease, which also spreads quickly due to its secretory nature. Therefore, its
diagnosis plays an important role in the disease assessment because of its mortal rate rounds
between 50-60% [163]. Recently, Narmani et al. developed an ultrasensitive and selective
fluorescence DNA biosensor based on AuNPs and magnetic NPs for the determination of the
bacteria’s O1 OmpW gene [164].
The gram-negative bacteria, Shigella, belongs to the Enterobacteriaceae family. Infected
people develop diverse symptoms including diarrhea, cramps, fever, and vomit. According to the
World Health Organization (WHO), the annual number of Shigella cases worldwide is
approximately 164.7 million with 1.1 million of those resulting in death, and the majority of them
involve young children under the age of 5 years old [165]. Research performed by Elahi et al.
discovered an early detection method of infectious Shigella. In this study, AuNPs-DNA probes were
hybridized with Spa gene sequence in order to create an optical genosensing system. This biosensor
makes the sample solution turns to purple in the absence of the complementary target, whereas the
solution remains red in the presence of the specific gene sequence [166]. On the other hand, Xiao et
al. covalently immobilized a DNA probe onto fiber-optic biosensors able to hybridize with a
fluorescently labeled complementary DNA. The obtained results were comparable to the ones
21
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
22
Sensors 2020, 20, x FOR PEER REVIEW 12 of 32
obtained by PCR, which suggests considering this method as an alternative for Shigella detection
[165].
The different approaches for biosensoring detection of pathogenic bacteria have been
successful and are currently being considered by many health governments and research
institutions, mainly because of their fast response, high-quality performance, and reliable results
[167–169].
4.2. Viral pathogen detection
Viral pathogen diagnosis is important for early and effective treatment in patients in order
to prevent outbreaks or pandemics. For that reason, biosensors are being widely employed for
making diagnosis easier, avoiding hard proteins or DNA identification techniques in specific virus
[170,171]. One of the most common and dangerous viral pathogens is the influenza virus because of
its ability to spread easily and constantly mutation. Hence, detection at early stages can be difficult
[172,173].
Hassanpour et al. developed a novel optical biosensor composed of pDNA bioconjugated
citrate capped AgNPs towards target sequences for ultrasensitive and selective Haemophilus
influenza detection in human biofluids [174]. This pathogen has also been detected through other
different biosensors, including the work reported by Jiang et al [148,174–176]. This paper describes
the development of a polydiacetylene sensitive biosensor using antibody detection for H5N1 (avian
influenza), in which the polydiacetylenes vesicles show a dramatic change in color from blue to red
upon the detection of the virus [176].
Other dangerous viruses that affect the population worldwide include ebolavirus, HIV, and
Hantavirus [177–179]. The first one is a negative strand-RNA virus that belongs to the Filoviridae
family and causes a deadly disease called Ebola. The infected people with this agent develop a
series of symptoms, where hemorrhagic fever is considered as fatal [180–182]. Currently, there is no
vaccine or specific treatment [183]. However, different studies have presented the development of
biosensors for detecting this pathogen [184]. Ilkhani et al. fabricated a novel electrochemical-based-
DNA biosensor through enzyme-amplified detection to improve the sensitivity and selectivity of
the device for the pathogen [185]. In addition to that, Baca et al. developed a biosensor that can
detect the virus within 10 minutes at the POC by using surface acoustic waves, showing potential to
detect it before symptoms onset [186].
On the other hand, HIV is a retrovirus that attacks a patient’s immune system, causing an
inability to resist many diseases, and culminating in death when the person is not under drug
control. Clinical treatments for HIV are crucial for reducing mortality, but early diagnosis saves
many lives as well and can decrease spread rates [187–189]. Shafiee et al. worked on a photonic
crystal biosensor to detect multiple HIV-1 subtypes (A, B, and D) upon binding of the biological
analyte with the biosensor [190]. In addition to that, Gong et al. prepared a nanocomposite of
polyaniline/graphene (PAN/GN) using reverse-phase polymerization for the development of an
electrical DNA-biosensor that showed great selectivity, and sensitivity for the detection of HIV-1
gene fragment [191].
Hantavirus is a cluster of viruses that are part of the Bunyaviridae family. The spread begins
through contact with liquids, food, or particles contaminated with rodent excreta. It causes
hemorrhagic fever, respiratory insufficiency, and heart failure within 2-7 days after getting infected
23
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
24
Sensors 2020, 20, x FOR PEER REVIEW 13 of 32
[192,193]. Regarding its detection, Gogola et al. have performed important research for the
development of biosensors [194,195]. In a first approach, they prepared an electrochemical
immunosensor based on chemical modification of the gold surface with the virus antigen/protein
[194]. In a second study, the research group designed a quick electrochemical biosensor based on
biochar (BC) as a carbonaceous platform for immunoassay applications due to its highly
functionalized surface for covalent binding with biomolecules [195]. Both studies developed
devices as promising and suitable tools for hantavirus clinical detection [194,195].
Furthermore, several bio-elements can be incorporated into a biosensor for virus detection
including markers, RNA, structural proteins, and enzymes from the viral pathogens [196].
5. COVID-19 pandemic
Currently, many viruses are being considered to have the capacity of causing future
pandemics. Different factors such as fast dissemination, a high transmission rate of new variants,
difficulties to develop efficient and sensible diagnostic techniques, as well as the lack of specific
vaccines and safe drugs for treatment, make them one of the major threats for mankind [197,198].
The most recent case is the COVID-19 announced as pandemic on March 13th, which is an infectious
disease with rapid human-to-human transmission caused by SARS-CoV-2. This pathogen belongs
to the positive-strand RNA viruses [199,200].
Like any other viral outbreak, an early diagnosis is fundamental for preventing an
uncontrollable spread of the disease. However, this pandemic has the particularity that more than
30% of the confirmed cases are asymptomatic, thus making it harder to control [200–202]. RT-PCR is
the most used suitable and reliable method for detecting SARS-CoV-2 infections until now.
Nevertheless, the technique is time-consuming, labor-intensive, and unavailable in remote settings
[203,204]. Although several other methods can be employed for that purpose, such as
immunological assays, thoracic imaging, portable X-rays, or amplification techniques, the pandemic
spread of COVID-19 demands to develop POC devices for rapid detection (Figure 3) [205–208].
25
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
26
Sensors 2020, 20, x FOR PEER REVIEW 14 of 32
Figure 3. POC for COVID-19.Reprinted with permission from Choi, J. et al. Development of Point-of-Care
Biosensors for COVID-19.Front Chem 8: 517. Copyright (2019) Frontiers in Chemistry [208].
Sheridan states that there are two types of rapid POC biosensors for COVID-19 detection.
In the first place, there is a nucleic acid test, which consists of detecting the virus in the patient’s
sputum, saliva, or nasal secretions [209,210]. The other type commonly employed is the antibody
test that is done through the analysis of collected blood samples five days after the initial infection,
which is when the immune response causes the production of IgM and IgG due to the presence of
the virus [211–213].
The industrial sector has developed some suitable POC biosensors for the qualitative
detection of SARS-CoV-2 IgM and IgG antibodies using samples as low as 10 µl of human serum,
whole blood, or finger prick, obtaining results within 10-15 minutes (Table 2) [214]. Many of these
rapid serological tests are paper-based biosensors that perform a colorimetric lateral flow
immunoassay. In this method, SARS-CoV-2 specific antigens are typically labeled with gold, and
bind the corresponding host antibodies, which migrate across an adhesive pad. As can be seen in
figure 4, anti-SARS-CoV-2 IgM antibodies interact with fixed anti-IgM secondary antibodies on the
M line, while IgG antibodies interact with anti-IgG antibodies on the G line. Therefore, M or G lines
only appear if the sample contains SARS-CoV-2 specific antibodies, otherwise, only the control line
(C) will be shown [215]. Although the use of serological tests to detect SARS-CoV-2 is still under
debate, these are foreseeing as crucial tools for the implementation or ceasing of lockdowns
established worldwide [216].
Table 2. FDA commercially authorized biosensors for SARS-CoV-2 detection [214].
27
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
28
Sensors 2020, 20, x FOR PEER REVIEW 15 of 32
Manufacturer Device Target
Clinical
combined
specificity
Clinical
combined
sensitivity
Abbott
SARS-CoV-2 IgG
chemilumininescent
microparticle immunoassay
(CMIA)
Nucleocapsid 99.9% 100%
Access Bio, Inc. CareStart COVID-19
IgM/IgG
Spike and
Nucleocapsid 98.9% 98.4%
Beijing Wantai
Biological
Pharmacy
Enterprise Co. Ltd.
Wantai SARS-CoV-2 Ab
rapid test Spike 98.8% 100%
Biohit Healthcare
(Hefei)
Biohit SARS-CoV-2 IgM/IgG
antibody test kit Nucleocapsid 95.0% 96.7%
Cellex
Cellex Qsars-CoV-2 IgG/IgM
rapid test lateral flow
immunoassay
Spike and
nucleocapsid 96.0% 93.8%
DiaSorin LIAISON SARS-CoV-2 S1/S2
IgG CMIA Spike 99.3% 97.6%
Hangzhou Biotest
Biotech
COVID-19 IgG/IgM rapid
test cassette Spike 100% 100%
Hangzhou Laihe
Biotech
LYHER novel coronavirus
(2019-nCoV) IgM/IgG
antibody combo test kit
(colloidal gold)
Spike 98.8% 100%
Healgen COVID-19 IgG/IgM rapid
test cassette Spike 97.5% 100%
Megna Health, Inc. Rapid COVID-19 IgM/IgG
combo test kit Nucleocapsid 95% 100%
Salofa Oy
Siena-Clarity COVIBLOCK
COVID-19 IgG/IgM Rapid
test cassette
Spike 98.8% 93.3%
Xiamen Biotime
Biotechnology Co.,
Ltd.
BIOTIME SARS-CoV-2
IgG/IgM rapid qualitative
test
Spike 96.2% 100%
CMIA: chemilumininescent microparticle immunoassay; COVID-19: coronavirus disease 2019.
29
484
30
Sensors 2020, 20, x FOR PEER REVIEW 16 of 32
Figure 4. COVID-19 rapid serological IgM/IgG test. Reprinted with permission from Ghaffari, A. et al.
COVID-19 Serological Test: How Well Do They Actually Perform? Diagnostics 10(7): 453. Copyright (2020)
MDPI [215].
Other research groups have developed Lab-on-a-Chip-based biosensors for SARS-CoV-2
detection [208,217]. This technology avoids the need for specialized personnel through the
integration of microfluidic components into a biosensor, allowing increasing their production, and
reducing the costs of the assay [218]. POC commercialized instruments based on this microfluidic
technology are having an important role in this pandemic, like ID NOW®, Filmarray®, GeneXpert®,
and RTisochip®[219].
Cell-based biosensors have also contributed to COVID-19 diagnosis. Mavrikou et al.
developed a biosensor based on membrane-engineered mammalian cells that possess the human
chimeric spike S1 antibody. The device can detect SARS-CoV-2 S1 spike protein selectively, where
the binding of the protein to the membrane-bound antibodies results in cellular bioelectric
properties modification measured by Bioelectric Recognition Assay. The LOD is 1 fg/ml and the
response time is about three minutes. In addition to that, the biosensor includes a portable read-out
device that can be operated by a smartphone [220].
Moreover, nano-biosensors have shown an outstanding potential to contribute to the fight
against COVID-19, providing holistic insights for developing ultrasensitive, cost-effective, and
rapid detection devices for mass production [221]. Advanced materials are the basis of nano-
enabled or integrated micro-and nano biosensing system technologies that can detect earlier the
virus, and even show good binding properties allowing them to inactive or destroy the pathogen
upon the application of an external stimulus [222].
Different research groups have developed carbon-based and graphene-based POC
biosensors [208,217]. Graphene is foreseeing to have a leading role in the attempt of fighting against
COVID-19. This low-cost material can be employed for virus detection since its sensitivity and
selectivity can be enhanced by modifying its hybrid structure (e.g., antibody-conjugated graphene
sheets) that allows tuning its optical and electrical features. Some graphene-based sensors that can
be explored for SARS-CoV-2 detection are photoluminescence, colorimetric, and SPR biosensors
31
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
32
Sensors 2020, 20, x FOR PEER REVIEW 17 of 32
[223,224]. Seo et al. employed the material for the development of a field-effect transistor (FET)-
based biosensor for detecting SARS-CoV-2 (Figure 5). In this case, graphene sheets from the FET
were coated with a specific antibody against the virus spike protein, which was successfully
detected at concentrations of 1 fg/ml in a phosphate-buffered saline medium. In addition, the
device was able to detect the virus in clinical samples, exhibiting a LOD of 2.42 × 102 copies/ml. The
fabricated biosensor is considered as a promising immunological diagnostic alternative for the
disease [225].
Figure 5.Schematic diagram of COVID-19 FET sensor operation procedure. Reprinted with
permission from Seo, G. et al. Rapid Detection of COVID-19 Causative Virus (SARS-CoV-2) in Human
Nasopharyngeal Swab Specimens Using Field-Effect Transistor-Based Biosensor.ACS Nano 14(4): 5135-5142.
Copyright (2020) ACS [225].
Additionally to FET, a review published by Cui et al. considers potential electrochemical
biosensor and surface-enhanced Raman scattering (SERS)-based biosensor as other suitable options
for diagnosis of COVID-19 [226]. Also, Murugan et al. designed two field-deployable/portable
plasmonic fiber-optic absorbance biosensor (P-FAB) device for rapid detection of the virus’ N-
protein directly from saliva. One of them was a labeled immunoassay, and the other one was label-
free. Both bioanalytical approaches using the highly sensitive P-FAB platform can be considered as
ideal alternatives for COVID-19 diagnosis within 15 minutes [227]. More recently, Zhu et al.
reported another diagnosis approach based on the development of a multiplex reverse transcription
loop-mediated isothermal amplification combined with NP-based lateral flow biosensor. The
method allowed the multiplex detection of the open reading frame 1a/b (ORF1ab) and the N-
protein within an hour, ensuring the sufficient sensitivity for the virus [228].
In another approach, Qiu et al. developed a dual-functional plasmonic biosensor that
combines the plasmonic photothermal (PPT) effect and LSPR sensing transduction. The device is
constituted by a two-dimensional gold nano island functionalized with complementary DNA
receptors that can selectively detect specific sequences from SARS-CoV-2 through nucleic acid
hybridization. In addition to that, PPT can increase the in situ hybridization temperature, which
allows differentiating between two similar gene sequences. This biosensor showed high sensitivity
with a lower LOD at 0.22 pM (i.e., 10-12 M) [229].
Although we have discussed several options for COVID-19 diagnosis, researchers are
working on novel diagnostic techniques that combine different approaches based on
nanotechnology and nanoscience, in order to obtain faster, reliable, and more accurate results that
33
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
34
Sensors 2020, 20, x FOR PEER REVIEW 18 of 32
allow accelerating life-saving decisions, and isolation of positive patients in an early stage to down-
regulate the virus spread [230,231].
6. Conclusions
Last few decades, viral and bacterial pathogens have become a real menace to human
safety. Their rapid identification must be considered as a priority task in order to prevent an
outbreak that represents a high risk of disruption of the healthcare system, and a disastrous socio-
economic impact. Scientists are performing intensive research for developing sensitive diagnostic
techniques and effective therapeutics. Although for many viruses and bacteria there is no vaccine or
pharmacological treatment, the development of a POC device for the rapid diagnosis of diseases
such as COVID-19, allows accelerating life-saving decisions, and isolation of positive patients in an
early stage. In this sense, biosensors and nano-biosensors are powerful measurement devices that
can make the detection process of important clinical bacteria and virus to be easy, quick, and
effective by sensing relevant parameter that can be related to infectious processes.
Author Contributions: Conceptualization, J.V.B. and R.C.S.; methodology, L.C.H., M.B.A., A.C.R., and J.V.B.;
investigation, L.C.H., M.B.A., and A.C.R.; resources, J.V.B.; writing—original draft preparation, L.C.H., M.B.A.,
A.C.R., and R.C.S.; writing—review and editing, L.C.H., and J.V.B.; visualization, M.B.A., A.C.R., and R.C.S.;
supervision, J.V.B.; project administration, J.V.B. All authors have read and agreed to the published version of
the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Clark LC, Lyons C. Electrode Systems for Continuous Monitoring in Cardiovascular Surgery. Ann N Y
Acad Sci. 1962;102(1):29–45.
2. Solaimuthu A, Vijayan AN, Murali P, Korrapati PS. Nano-biosensors and their relevance in tissue
engineering. Curr Opin Biomed Eng. 2020;13:84–93. Doi:10.1016/j.cobme.2019.12.005
3. Metkar SK, Girigoswami K. Diagnostic biosensors in medicine A review. Biocatal Agric
Biotechnol.2019;17:271–83. Doi:10.1016/j.bcab.2018.11.029
4. Lakshmipriya T, Gopinath SCB. 1 - An Introduction to Biosensors and Biomolecules. In: Gopinath SCB,
Lakshmipriya T. Nanobiosensors for Biomolecular Targeting. Elsevier; 2019. p. 1–21. Doi:10.1016/B978-0-
12-813900-4.00001-4
5. Scholten K, Meng E. A review of implantable biosensors for closed-loop glucose control and other drug
delivery applications. Int J Pharm. 2018;544(2):319–34. Doi: 10.1016/j.ijpharm.2018.02.022
6. Yazdi MK, Zarrintaj P, Bagheri B, Kim YC, Ganjali MR, Saeb MR. 32 - Nanotechnology-based biosensors
in drug delivery. In: Mozafari M. Nanoengineered Biomaterials for Advanced Drug Delivery. Woodhead
Publishing Series in Biomaterials.2020 .p. 767–79. Doi:10.1016/b978-0-08-102985-5.00032-2
7. Griesche C, Baeumner AJ. Biosensors to support sustainable agriculture and food safety. TrAC Trends
Anal Chem. 2020;128:115906. Doi: 10.1016/j.trac.2020.115906
8. Pandey A, Gurbuz Y, Ozguz V, Niazi JH, Qureshi A. Graphene-interfaced electrical biosensor for label-
free and sensitive detection of foodborne pathogenic E. coli O157:H7. Biosens Bioelectron. 2017;91:225–
31.doi: 10.1016/j.bios.2016.12.041
9. Cesewski E, Johnson BN. Electrochemical biosensors for pathogen detection. Biosens Bioelectron. 2020 Jul
1;159:112214. Doi: 10.1016/j.bios.2020.112214
35
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
36
Sensors 2020, 20, x FOR PEER REVIEW 19 of 32
10. Guliy OI, Zaitsev BD, Larionova OS, Borodina IA.Virus Detection Methods and Biosensor
Technologies.Biophysics. 2019;64(6):890–7. Doi: 10.1134/S0006350919060095
11. Farooq U, Yang Q, Ullah MW, Wang S. Bacterial biosensing: Recent advances in phage-based bioassays
and biosensors. Biosens Bioelectron. 2018;118:204–16. Doi: 10.1016/j.bios.2018.07.058
12. Cheng MS, Ho JS, Tan CH, Wong JPS, Ng LC, Toh C-S. Development of an electrochemical membrane-
based nanobiosensor for ultrasensitive detection of dengue virus.Anal Chim Acta. 2012;725:74–80. Doi:
10.1016/j.aca.2012.03.017
13. Sharma A, Sharma N, Kumari A, Lee H-J, Kim T, Tripathi KM. Nano-carbon based sensors for bacterial
detection and discrimination in clinical diagnosis: A junction between material science and biology. Appl
Mater Today. 2020;18:100467. Doi: 10.1016/j.apmt.2019.100467
14. Rai M, Gade A, Gaikwad S, Marcato PD, Durán N. Biomedical applications of nanobiosensors: the state-
of-the-art. J Braz Chem Soc. 2012;23(1):14–24. Doi:10.1590/S0103-50532012000100004
15. Zhao VXT, Wong TI, Zheng XT, Tan YN, Zhou X. Colorimetric biosensors for point-of-care virus
detections. Mater Sci Energy Technol. 2020;3:237–49. Doi: 10.1016/j.mset.2019.10.002
16. Stringer RC, Schommer S, Hoehn D, Grant SA. Development of an optical biosensor using gold
nanoparticles and quantum dots for the detection of Porcine Reproductive and Respiratory Syndrome
Virus. Sens Actuators B Chem. 2008;134(2):427–31. Doi: 10.1016/j.snb.2008.05.018
17. Dell’Atti D, Zavaglia M, Tombelli S, Bertacca G, Cavazzana AO, Bevilacqua G, et al. Development of
combined DNA-based piezoelectric biosensors for the simultaneous detection and genotyping of high
risk Human Papilloma Virus strains. Clin Chim Acta. 2007;383(1):140–6. Doi: 10.1016/j.cca.2007.05.009
18. Du K, Cai H, Park M, Wall TA, Stott MA, Alfson KJ, et al. Multiplexed efficient on-chip sample
preparation and sensitive amplification-free detection of Ebola virus. Biosens Bioelectron. 2017;91:489–96.
Doi: 10.1016/j.bios.2016.12.071
19. Hu Y, Li H, Li J. A novel electrochemical biosensor for HIV-related DNA detection based on toehold
strand displacement reaction and cruciform DNA crystal. J Electroanal Chem. 2018;822:66–72. Doi:
10.1016/j.jelechem.2018.05.011
20. Seo SE, Tabei F, Park SJ, Askarian B, Kim KH, Moallem G, et al. Smartphone with optical, physical, and
electrochemical nanobiosensors. J Ind Eng Chem. 2019;77:1–11. Doi: 10.1016/j.jiec.2019.04.037
21. Monošík R, Stred’anský M, Šturdík E. Application of Electrochemical Biosensors in Clinical Diagnosis. J
Clin Lab Anal. 2012;26(1):22–34.
22. Mobed A, Baradaran B, Guardia M de la, Agazadeh M, Hasanzadeh M, Rezaee MA, et al. Advances in
detection of fastidious bacteria: From microscopic observation to molecular biosensors. TrAC Trends
Anal Chem. 2019;113:157–71. Doi:10.1016/j.trac.2019.02.012
23. Petrosova A, Konry T, Cosnier S, Trakht I, Lutwama J, Rwaguma E, et al. Development of a highly
sensitive, field operable biosensor for serological studies of Ebola virus in central Africa. Sens Actuators B
Chem. 2007;122(2):578–86. Doi:10.1016/j.snb.2006.07.005
24. Encarnação JM, Rosa L, Rodrigues R, Pedro L, da Silva FA, Gonçalves J, et al. Piezoelectric biosensors for
biorecognition analysis: Application to the kinetic study of HIV-1 Vif protein binding to recombinant
antibodies. J Biotechnol. 2007;132(2):142–8. Doi: 10.1016/j.jbiotec.2007.04.010
25. Sitdikov RA, Wilkins ES, Yates T, Hjelle B. Detection of Hantavirus Using a New Miniaturized Biosensor
Device. J Appl Res. 2007;7(1):22.
26. Ionescu RE. Biosensor Platforms for Rapid Detection of E. coli Bacteria. In: Samie A. Recent Advances on
Physiology, Pathogenesis and Biotechnological Applications. 2017. IntechOpen. Doi:10.5772/67392
37
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
38
Sensors 2020, 20, x FOR PEER REVIEW 20 of 32
27. Malvano F, Pilloton R, Albanese D. A novel impedimetric biosensor based on the antimicrobial activity of
the peptide nisin for the detection of Salmonella spp. Food Chem. 2020;325:126868. Doi:
10.1016/j.foodchem.2020.126868
28. Chao J, Zhu D, Zhang Y, Wang L, Fan C. DNA nanotechnology-enabled biosensors. Biosens Bioelectron.
2016;76:68–79. Doi: 10.1016/j.bios.2015.07.007
29. Soni A, Surana RK, Jha SK. Smartphone based optical biosensor for the detection of urea in saliva. Sens
Actuators B Chem. 2018;269:346–53. Doi: 10.1016/j.snb.2018.04.108
30. Zhang H, Xue L, Huang F, Wang S, Wang L, Liu N, et al. A capillary biosensor for rapid detection of
Salmonella using Fe-nanocluster amplification and smart phone imaging.Biosens Bioelectron.
2019;127:142–9. Doi: 10.1016/j.bios.2018.11.042
31. Roda A, Michelini E, Zangheri M, Di Fusco M, Calabria D, Simoni P. Smartphone-based biosensors: A
critical review and perspectives. TrAC Trends Anal Chem. 2016;79:317–25. doi: 10.1016/j.trac.2015.10.019
32. Choi C. Integrated nanobiosensor technology for biomedical application.Vol. 1, Nanobiosensors in
Disease Diagnosis. 2012;1:1-4. Doi: 10.2147/NDD.S26422
33. Srinivasan B, Tung S. Development and Applications of Portable Biosensors. SLAS TECH. 2015;20(4):365-
389.Doi:10.1177/2211068215581349
34. Mehrotra P. Biosensors and their applications - A review. J Oral Biol Craniofacial Res. 2016;6(2):153–9.
Doi: 10.1016/j.jobcr.2015.12.002
35. Krejcova L, Michalek P, Rodrigo MM, Heger Z, Krizkova S, Vaculovicova M, et al. Nanoscale virus
biosensors: state of the art. Nanobiosensors in Disease Diagnosis.2015;4:47–66. Doi: 10.2147/NDD.S56771
36. Malon RSP, Sadir S, Balakrishnan M, Córcoles EP. Saliva-Based Biosensors: Noninvasive Monitoring Tool
for Clinical Diagnostics. BioMed Res Int. 2014; 962903. Doi: 10.1155/2014/962903
37. Zhang W, Du Y, Wang ML. Noninvasive glucose monitoring using saliva nano-biosensor. Sens Bio-Sens
Res. 2015;4:23–9. Doi: 10.1016/j.sbsr.2015.02.002
38. Mishra RK, Rajakumari R. Chapter 1 - Nanobiosensors for Biomedical Application: Present and Future
Prospects. In: Mohapatra SS, Ranjan S, Dasgupta N, Mishra RK, Thomas S. Characterization and Biology
of Nanomaterials for Drug Delivery. Elsevier; 2019. p. 1–23. Doi: 10.1016/B978-0-12-814031-4.00001-5
39. Aiello V, Fichera M, Giannazzo F, Libertino S, Scandurra A, Reins M, et al. Fabrication and
characterization of the sensing element for glucose biosensor applications. In: Sensors and Microsystems.
WORLD SCIENTIFIC; 2008. Doi:10.1142/9789812833594_0001
40. Selvarajan S, Alluri NR, Chandrasekhar A, Kim SJ. Unconventional active biosensor made of piezoelectric
BaTiO3 nanoparticles for biomolecule detection. Sens Actuators B Chem. 2017;253:1180–7. Doi:
10.1016/j.snb.2017.07.159
41. Bhatnagar I, Mahato K, Ealla KKR, Asthana A, Chandra P. Chitosan stabilized gold nanoparticle
mediated self-assembled gliP nanobiosensor for diagnosis of Invasive Aspergillosis. Int J Biol Macromol.
2018;110:449–56. Doi: 10.1016/j.ijbiomac.2017.12.084
42. Riu J, Giussani B. Electrochemical biosensors for the detection of pathogenic bacteria in food. TrAC
Trends Anal Chem. 2020;126:115863. Doi: 10.1016/j.trac.2020.115863
43. Gupta R, Raza N, Bhardwaj SK, Vikrant K, Kim K-H, Bhardwaj N. Advances in nanomaterial-based
electrochemical biosensors for the detection of microbial toxins, pathogenic bacteria in food matrices. J
Hazard Mater. 2020;123379. Doi: 10.1016/j.jhazmat.2020.123379
44. Zhao W, Xing Y, Lin Y, Gao Y, Wu M, Xu J. Monolayer graphene chemiresistive biosensor for rapid
bacteria detection in a microchannel. Sens Actuators Rep. 2020;2(1):100004. Doi: 10.1016/j.snr.2020.100004
39
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
40
Sensors 2020, 20, x FOR PEER REVIEW 21 of 32
45. Samanman S, Kanatharana P, Chotigeat W, Deachamag P, Thavarungkul P. Highly sensitive capacitive
biosensor for detecting white spot syndrome virus in shrimp pond water. J Virol Methods.
2011;173(1):75–84. Doi: 10.1016/j.jviromet.2011.01.010
46. Huang Y, Xu J, Liu J, Wang X, Chen B. Disease-Related Detection with Electrochemical Biosensors: A
Review. Sensors. 2017;17(10). Doi: 10.3390/s17102375
47. Fu Z, Lu YC, Lai JJ. Recent Advances in Biosensors for Nucleic Acid and Exosome Detection. Chonnam
Med J. 2019;55(2):86–98. Doi: 10.4068/cmj.2019.55.2.86
48. Wang P, Menzies NW, Lombi E, Sekine R, Blamey FPC, Hernandez-Soriano MC, et al. Silver sulfide
nanoparticles (Ag2S-NPs) are taken up by plants and are phytotoxic. Nanotoxicology. 2015;9(8):1041–9.
Doi:10.3109/17435390.2014.999139
49. García-Aljaro C, Cella LN, Shirale DJ, Park M, Muñoz FJ, Yates MV, et al. Carbon nanotubes-based
chemiresistive biosensors for detection of microorganisms. Biosens Bioelectron. 2010;26(4):1437–41. Doi:
10.1016/j.bios.2010.07.077
50. Nosrati R, Dehghani S, Karimi B, Yousefi M, Taghdisi SM, Abnous K, et al. Siderophore-based biosensors
and nanosensors; new approach on the development of diagnostic systems. Biosens Bioelectron.
2018;117:1–14. Doi: 10.1016/j.bios.2018.05.057
51. Hanif A, Farooq R, Rehman MU, Khan R, Majid S, Ganaie MA. Aptamer based nanobiosensors:
Promising healthcare devices. Saudi Pharm J. 2019;27(3):312–9. Doi: 10.1016/j.jsps.2018.11.013
52. Lei Y, Chen W, Mulchandani A. Microbial biosensors. Anal Chim Acta. 2006;568(1–2):200–10. Doi:
10.1016/j.aca.2005.11.065
53. Socorro-Leránoz AB, Santano D, Del Villar I, Matias IR. Trends in the design of wavelength-based optical
fibre biosensors (2008–2018).Biosens Bioelectron X. 2019;1:100015. Doi: 10.1016/j.biosx.2019.100015
54. Tran LD, Nguyen BH, Van Hieu N, Tran HV, Nguyen HL, Nguyen PX. Electrochemical detection of short
HIV sequences on chitosan/Fe3O4 nanoparticle based screen printed electrodes. Mater Sci Eng C.
2011;31(2):477–85. Doi: 10.1016/j.msec.2010.11.007
55. Lazerges M, Bedioui F. Analysis of the evolution of the detection limits of electrochemical DNA
biosensors. Anal Bioanal Chem. 2013;405(11):3705–14. Doi: 10.1007/s00216-012-6672-5
56. Solanki PR, Patel MK, Kaushik A, Pandey MK, Kotnala RK, Malhotra BD. Sol-Gel Derived
Nanostructured Metal Oxide Platform for Bacterial Detection. Electroanalysis. 2011;23(11):2699–708. Doi:
10.1002/elan.201100351
57. Shabaninejad Z, Yousefi F, Movahedpour A, Ghasemi Y, Dokanehiifard S, Rezaei S, et al.
Electrochemical-based biosensors for microRNA detection: Nanotechnology comes into view. Anal
Biochem. 2019;581:113349. Doi: 10.1016/j.ab.2019.113349
58. Leung A, Shankar PM, Mutharasan R. A review of fiber-optic biosensors. Sens Actuators B Chem.
2007;125(2):688–703. Doi: 10.1016/j.snb.2007.03.010
59. Benito-Peña E, Valdés MG, Glahn-Martínez B, Moreno-Bondi MC. Fluorescence based fiber optic and
planar waveguide biosensors. A review.Anal Chim Acta. 2016;943:17–40. Doi: 10.1016/j.aca.2016.08.049
60. Zhai Q, Cheng W. Soft and stretchable electrochemical biosensors. Mater Today Nano. 2019;7:100041. Doi:
10.1016/j.mtnano.2019.100041
61. Skládal P. Piezoelectric biosensors. TrAC Trends Anal Chem. 2016;79:127–33. Doi:
10.1016/j.trac.2015.12.009
62. Sawant SN. 13 - Development of Biosensors From Biopolymer Composites. In: Sadasivuni KK,
Ponnamma D, Kim J, Cabibihan JJ, AlMaadeed MA. Biopolymer Composites in Electronics.Elsevier; 2017.
p. 353–83. Doi: 10.1016/B978-0-12-809261-3.00013-9
41
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
42
Sensors 2020, 20, x FOR PEER REVIEW 22 of 32
63. Bhalla N, Jolly P, Formisano N, Estrela P. Introduction to biosensors. Essays Biochem. 2016;60(1):1–8. Doi:
10.1042/EBC20150001
64. Khansili N, Rattu G, Krishna PM. Label-free optical biosensors for food and biological sensor
applications. Sens Actuators B Chem. 2018;265:35–49. Doi:10.1016/j.snb.2018.03.004
65. Konopsky VN, Alieva EV. Imaging biosensor based on planar optical waveguide. Opt Laser Technol.
2019;115:171–5. Doi: 10.1016/j.optlastec.2019.02.034
66. Yin M, Gu B, An QF, Yang C, Guan YL, Yong KT. Recent development of fiber-optic chemical sensors and
biosensors: Mechanisms, materials, micro/nano-fabrications and applications. Coord Chem Rev.
2018;376:348–92. Doi: 10.1016/j.ccr.2018.08.001
67. Chen Y, Liu J, Yang Z, Wilkinson JS, Zhou X. Optical biosensors based on refractometric sensing schemes:
A review. Biosens Bioelectron. 2019;144:111693. Doi:10.1016/j.bios.2019.111693
68. Vidal P, Martinez M, Hernandez C, Adhikari AR, Mao Y, Materon L, et al. Development of chromatic
biosensor for quick bacterial detection based on polyvinyl butyrate-polydiacetylene nonwoven fiber
composites. Eur Polym J. 2019;121:109284. Doi: 10.1016/j.eurpolymj.2019.109284
69. Jeong W, Choi S-H, Lee H, Lim Y. A fluorescent supramolecular biosensor for bacterial detection via
binding-induced changes in coiled-coil molecular assembly. Sens Actuators B Chem. 2019;290:93–9. Doi:
10.1016/j.snb.2019.03.112
70. Ahmadi H, Heidarzadeh H, Taghipour A, Rostami A, Baghban H, Dolatyari M, et al. Evaluation of single
virus detection through optical biosensor based on microsphere resonator. Optik. 2014;125(14):3599–602.
Doi: 10.1016/j.ijleo.2014.01.087
71. Blair EO, Corrigan DK. A review of microfabricated electrochemical biosensors for DNA detection.
Biosens Bioelectron. 2019;134:57–67. Doi: 10.1016/j.bios.2019.03.055
72. Hammond JL, Formisano N, Estrela P, Carrara S, Tkac J. Electrochemical biosensors and nanobiosensors.
Essays Biochem. 2016;60(1):69–80. Doi: 10.1042/EBC20150008
73. Wang H, Li H, Huang Y, Xiong M, Wang F, Li C. A label-free electrochemical biosensor for highly
sensitive detection of gliotoxin based on DNA nanostructure/MXene nanocomplexes. Biosens
Bioelectron. 2019;142:111531. Doi: 10.1016/j.bios.2019.111531
74. Mathelié M, Cohen T, Gammoudi I, Martin A, Béven L, Delville MH, et al. Silica nanoparticles-assisted
electrochemical biosensor for the rapid, sensitive and specific detection of Escherichia coli. Sens Actuators
B Chem. 2019;292:314–20. Doi: 10.1016/j.snb.2019.03.144
75. Baek SH, Kim MW, Park CY, Choi CS, Kailasa SK, Park JP, et al. Development of a rapid and sensitive
electrochemical biosensor for detection of human norovirus via novel specific binding peptides. Biosens
Bioelectron. 2019;123:223–9. Doi: 10.1016/j.bios.2018.08.064
76. Pohanka M. Overview of Piezoelectric Biosensors, Immunosensors and DNA Sensors and Their
Applications.Materials. 2018;11(3). Doi: 10.3390/ma11030448
77. Tombelli S, Minunni M, Santucci A, Spiriti MM, Mascini M. A DNA-based piezoelectric biosensor:
Strategies for coupling nucleic acids to piezoelectric devices. Talanta. 2006;68(3):806–12. Doi:
10.1016/j.talanta.2005.06.007
78. Tombelli S. 2 - Piezoelectric biosensors for medical applications. In: Higson S. Biosensors for Medical
Applications. Woodhead Publishing; 2012. p. 41–64. Doi: 10.1533/9780857097187.1.41
79. Fu YQ, Luo JK, Nguyen NT, Walton AJ, Flewitt AJ, Zu XT, et al. Advances in piezoelectric thin films for
acoustic biosensors, acoustofluidics and lab-on-chip applications. Prog Mater Sci. 2017;89:31–91. Doi:
10.1016/j.pmatsci.2017.04.006
43
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
44
Sensors 2020, 20, x FOR PEER REVIEW 23 of 32
80. Guo X, Lin C-S, Chen S-H, Ye R, Wu VCH. A piezoelectric immunosensor for specific capture and
enrichment of viable pathogens by quartz crystal microbalance sensor, followed by detection with
antibody-functionalized gold nanoparticles. Biosens Bioelectron. 2012;38(1):177–83. Doi:
10.1016/j.bios.2012.05.024
81. Shamsipur M, Nasirian V, Mansouri K, Barati A, Veisi-Raygani A, Kashanian S. A highly sensitive
quantum dots-DNA nanobiosensor based on fluorescence resonance energy transfer for rapid detection
of nanomolar amounts of human papillomavirus 18.J Pharm Biomed Anal. 2017;136:140–7. Doi:
10.1016/j.jpba.2017.01.002
82. Golichenari B, Nosrati R, Farokhi A, Abnous K, Vaziri F, Behravan J. Nano-biosensing approaches on
tuberculosis: Defy of aptamers. Biosens Bioelectron. 2018;117:319–31. Doi: 10.1016/j.bios.2018.06.025
83. Malik P, Katyal V, Malik V, Asatkar A, Inwati G, Mukherjee TK. Nanobiosensors: Concepts and
Variations. ISRN Nanomaterials. 2013;327435. Doi: 10.1155/2013/327435
84. Park CS, Lee C, Kwon OS. Conducting Polymer Based Nanobiosensors. Polymers.2016;8(7):249. Doi:
10.3390/polym8070249
85. Chamorro A, Merkoci A. Nanobiosensors in diagnostics. Nanobiomedicine. 2016;3. Doi:
10.1177/1849543516663574
86. Harvey JD, Baker HA, Ortiz MV, Kentsis A, Heller DA. HIV Detection via a Carbon Nanotube RNA
Sensor. ACS Sens. 2019;4(5):1236–44. Doi: 10.1021/acssensors.9b00025
87. Lee J, Takemura K, Park EY. Plasmonic Nanomaterial-Based Optical Biosensing Platforms for Virus
Detection.Sensors. 2017;17(10):2332. Doi: 10.3390/s17102332
88. Mokhtarzadeh A, Eivazzadeh R, Pashazadeh P, Hejazi M, Gharaatifar N, Hasanzadeh M, et al.
Nanomaterial-based biosensors for detection of pathogenic virus. TrAC Trends Anal Chem. 2017;97:445–
57. Doi: 10.1016/j.trac.2017.10.005
89. Lin Z, Wu G, Zhao L, Lai KW. Carbon Nanomaterial-Based Biosensors: A Review of Design and
Applications. IEEE Nanotechnol Mag. 2019;13(5):4–14. Doi: 10.1109/MNANO.2019.2927774
90. Mansuriya BD, Altintas Z. Applications of Graphene Quantum Dots in Biomedical Sensors. Sensors.
2020;20(4). Doi: 10.3390/s20041072
91. Pirzada M, Altintas Z. Nanomaterials for Healthcare Biosensing Applications. Sensors. 2019;19(23):5311.
Doi: 10.3390/s19235311
92. Holzinger M, Le Goff A, Cosnier S. Nanomaterials for biosensing applications: a review. Front Chem.
2014;2(63):25221775. Doi: 10.3389/fchem.2014.00063
93. Campuzano S, Yáñez P, Pingarrón JM. Carbon Dots and Graphene Quantum Dots in Electrochemical
Biosensing.Nanomaterials. 2019;9(4). Doi: 10.3390/nano9040634
94. Jyoti A, Tomar RS. Detection of pathogenic bacteria using nanobiosensors. Environ Chem Lett.
2017;15(1):1–6. Doi: 10.1007/s10311-016-0594-y
95. Adegoke O, Seo M, Kato T, Kawahito S, Park E. An ultrasensitive SiO2-encapsulated alloyed CdZnSeS
quantum dot-molecular beacon nanobiosensor for norovirus. Biosens Bioelectron. 2016;86:135–42. Doi:
10.1016/j.bios.2016.06.027
96. Younis S, Taj A, Zia R, Hayat H, Shaheen A, Awan FR, et al. Chapter 19 - Nanosensors for the detection of
viruses. In: Han B, Tomer VK, Nguyen TA, Farmani A, Kumar Singh P. Nanosensors for Smart Cities.
Elsevier; 2020. p. 327–38. Doi: 10.1016/B978-0-12-819870-4.00018-9
97. Khan I, Saeed K, Khan I. Nanoparticles: Properties, applications and toxicities. Arab J Chem.
2019;12(7):908–31. Doi: 10.1016/j.arabjc.2017.05.011
45
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
46
Sensors 2020, 20, x FOR PEER REVIEW 24 of 32
98. Saleh TA, Gupta VK. Chapter 4 - Synthesis, Classification, and Properties of Nanomaterials. In: Saleh TA,
Gupta VK.Nanomaterial and Polymer Membranes.Elsevier; 2016.p. 83–133. Doi:10.1016/B978-0-12-
804703-3.00004-8
99. Guleria A, Neogy S, Raorane BS, Adhikari S. Room temperature ionic liquid assisted rapid synthesis of
amorphous Se nanoparticles: Their prolonged stabilization and antioxidant studies. Mater Chem Phys.
2020;253:123369. Doi: 10.1016/j.matchemphys.2020.123369
100. Sudha PN, Sangeetha K, Vijayalakshmi K, Barhoum A. Chapter 12 - Nanomaterials history, classification,
unique properties, production and market. In: Barhoum A, Makhlouf AS. Emerging Applications of
Nanoparticles and Architecture Nanostructures.Elsevier; 2018. p. 341–84. Doi: 10.1016/B978-0-323-51254-
1.00012-9
101. Al-Qadi S, Remuñan C. Nanopartículas metálicas: oro. In: Vila JL. Nanotecnología Farmacéutica:
Realidades y Posibilidades Farmacoterapéuticas. Real Academia Nacional de Farmacia; 2008.
102. Gómez P, Puente A, Castro A, Santos do Nascimento LA, Balu AM, Luque R, et al. Nanomaterials and
catalysis for green chemistry. Curr Opin Green Sustain Chem. 2020;24:48–55. Doi:
10.1016/j.cogsc.2020.03.001
103. Kalimuthu K, Cha BS, Kim S, Park KS. Eco-friendly synthesis and biomedical applications of gold
nanoparticles: A review. Microchem J. 2020;152:104296. Doi: 10.1016/j.microc.2019.104296
104. Calderón B, Johnson ME, Montoro AR, Murphy KE, Winchester MR, Vega JR. Silver Nanoparticles:
Technological Advances, Societal Impacts, and Metrological Challenges. Front Chem. 2017;5:6. Doi:
10.3389/fchem.2017.00006
105. Solano V, Vega J. Gold, silver, copper and silicone hybrid nanostructure cytotoxicity. Int J Rec Sci Res.
2017;8(2):15478-15486.
106. Bhardwaj N, Bhardwaj SK, Bhatt D, Lim DK, Kim K-H, Deep A. Optical detection of waterborne
pathogens using nanomaterials. TrAC Trends Anal Chem. 2019;113:280–300. Doi:
10.1016/j.trac.2019.02.019
107. Elahi N, Kamali M, Baghersad MH, Amini B. A fluorescence Nano-biosensors immobilization on Iron
(MNPs) and gold (AuNPs) nanoparticles for detection of Shigella spp. Mater Sci Eng C. 2019;105:110113.
Doi: 10.1016/j.msec.2019.110113
108. Takemura K, Adegoke O, Takahashi N, Kato T, Li TC, Kitamoto N, et al. Versatility of a localized surface
plasmon resonance-based gold nanoparticle-alloyed quantum dot nanobiosensor for
immunofluorescence detection of viruses. Biosens Bioelectron. 2017;89:998–1005. Doi:
10.1016/j.bios.2016.10.045
109. Suvarnaphaet P, Pechprasarn S. Graphene-Based Materials for Biosensors: A Review. Sensors.
2017;17(10). Doi: 10.3390/s17102161
110. Safardoust H, Salavati M, Amiri O, Hassanpour M. Preparation of highly luminescent nitrogen doped
graphene quantum dots and their application as a probe for detection of Staphylococcus aureus and E.
coli. J Mol Liq. 2017;241:1114–9. Doi: 10.1016/j.molliq.2017.06.106
111. Mat MH, Abdullah J, Yusof NA, Sulaiman Y, Wasoh H, Noh MF, et al. PNA biosensor based on reduced
graphene oxide/water soluble quantum dots for the detection of Mycobacterium tuberculosis. Sens
Actuators B Chem. 2017;241:1024–34. Doi: 10.1016/j.snb.2016.10.045
112. Duan N, Wu S, Dai S, Miao T, Chen J, Wang Z. Simultaneous detection of pathogenic bacteria using an
aptamer based biosensor and dual fluorescence resonance energy transfer from quantum dots to carbon
nanoparticles. Microchim Acta. 2015;182(5–6):917–23. Doi:10.1007/s00604-014-1406-3
47
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
48
Sensors 2020, 20, x FOR PEER REVIEW 25 of 32
113. Castillo L, Vargas R, Pacheco J, Vega J. Electrospun nanofibers: A nanotechnological approach for drug
delivery and dissolution optimization in poorly water-soluble drugs. ADMET and DMPK. 2020;8. Doi:
10.5599/admet.844
114. Schaub N. Electrospun fibers: a guiding scaffold for research and regeneration of the spinal cord. Neu
Reg Res.2016;11(11):1764-1765. Doi: 10.4103/1673-5374.194719
115. Ding Y, Li W, Zhang F, Liu Z, Ezazi NZ, Liu D, et al. Electrospun Fibrous Architectures for Drug
Delivery, Tissue Engineering and Cancer Therapy. Adv Funct Mater. 2019;29(2):1802852. Doi:
10.1002/adfm.201802852
116. Potrč T, Baumgartner S, Roškar R, Planinšek O, Lavrič Z, Kristl J, et al. Electrospun polycaprolactone
nanofibers as a potential oromucosal delivery system for poorly water-soluble drugs. Eur J Pharm Sci.
2015;75:101–13. Doi: 10.1016/j.ejps.2015.04.004
117. Zhang S, Jia Z, Liu T, Wei G, Su Z. Electrospinning Nanoparticles-Based Materials Interfaces for Sensor
Applications. Sensors. 2019;19(18):3977. Doi: 10.3390/s19183977
118. Aliheidari N, Aliahmad N, Agarwal M, Dalir H. Electrospun Nanofibers for Label-Free Sensor
Applications. Sensors. 2019;19(16):3587. Doi: 10.3390/s19163587
119. Maslakci NN, Ulusoy S, Oksuz AU. Investigation of the effects of plasma-treated chitosan electrospun
fibers onto biofilm formation. Sens Actuators B Chem. 2017;246:887–95. Doi: 10.1016/j.snb.2017.02.089
120. Asmatulu R, Khan WS. Chapter 9 - Electrospun nanofibers for nanosensor and biosensor applications. In:
Asmatulu R, Khan WS. Synthesis and Applications of Electrospun Nanofibers.Elsevier; 2019. p. 175–96.
Doi: 10.1016/B978-0-12-813914-1.00009-2
121. Tang L, Xie X, Zhou Y, Zeng G, Tang J, Wu Y, et al. A reusable electrochemical biosensor for highly
sensitive detection of mercury ions with an anionic intercalator supported on ordered mesoporous
carbon/self-doped polyaniline nanofibers platform. Biochem Eng J. 2017;117:7–14. Doi:
10.1016/j.bej.2016.09.011
122. Kafi AK, Wali Q, Jose R, Biswas TK, Yusoff MM. A glassy carbon electrode modified with SnO2
nanofibers, polyaniline and hemoglobin for improved amperometric sensing of hydrogen peroxide.
Microchim Acta. 2017;184(11):4443–50. Doi:10.1007/s00604-017-2479-6
123. Sapountzi E, Braiek M, Chateaux JF, Jaffrezic N, Lagarde F. Recent Advances in Electrospun Nanofiber
Interfaces for Biosensing Devices.Sensors. 2017;17(8):1887. Doi: 10.3390/s17081887
124. Supraja P, Tripathy S, Krishna SR, Singh V, Singh SG. Label free, electrochemical detection of atrazine
using electrospun Mn2O3 nanofibers: Towards ultrasensitive small molecule detection. Sens Actuators B
Chem. 2019;285:317–25. Doi: 10.1016/j.snb.2019.01.060
125. Sundera S, Mohamed MS, Perumal V, Mohamed MS, Mohamed NM. 11 - Electrospun Nanofibers for
Biosensing Applications. In: Gopinath SC, Lakshmipriya T. Nanobiosensors for Biomolecular Targeting.
Elsevier; 2019. p. 253–67. Doi: 10.1016/B978-0-12-813900-4.00011-7
126. Maftoonazad N, Ramaswamy H. Design and testing of an electrospun nanofiber mat as a pH biosensor
and monitor the pH associated quality in fresh date fruit (Rutab). Polym Test. 2019;75:76–84. Doi:
10.1016/j.polymertesting.2019.01.011
127. Wang X, Wang Y, Shan Y, Jiang M, Gong M, Jin X, et al. An electrochemiluminescence biosensor for
detection of CdkN2A/p16 anti-oncogene based on functional electrospun nanofibers and core-shell
luminescent composite nanoparticles.Talanta. 2018;187:179–87. Doi: 10.1016/j.talanta.2018.05.033
128. Matlock L, Coon B, Pitner CL, Frey MW, Baeumner AJ. Functionalized electrospun poly(vinyl alcohol)
nanofibers for on-chip concentration of E. coli cells. Anal Bioanal Chem. 2016;408(5):1327–34. Doi:
10.1007/s00216-015-9112-5
49
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
50
Sensors 2020, 20, x FOR PEER REVIEW 26 of 32
129. Luo Y, Nartker S, Miller H, Hochhalter D, Wiederoder M, Wiederoder S, et al. Surface functionalization
of electrospun nanofibers for detecting E. coli O157:H7 and BVDV cells in a direct-charge transfer
biosensor. Biosens Bioelectron. 2010;26(4):1612–7. Doi: 10.1016/j.bios.2010.08.028
130. Quirós J, Amaral AJ, Pasparakis G, Williams GR, Rosal R. Electrospun boronic acid-containing polymer
membranes as fluorescent sensors for bacteria detection. React Funct Polym. 2017;121:23–31. Doi:
10.1016/j.reactfunctpolym.2017.10.007
131. Tripathy S, Krishna SR, Singh V, Swaminathan S, Singh SG. Electrospun manganese (III) oxide nanofiber
based electrochemical DNA-nanobiosensor for zeptomolar detection of dengue consensus primer.
Biosens Bioelectron. 2017;90:378–87. Doi: 10.1016/j.bios.2016.12.008
132. Liu Y, Hao M, Chen Z, Liu L, Liu Y, Yang W, et al. A review on recent advances in application of
electrospun nanofiber materials as biosensors.Curr Opin Biomed Eng. 2020;13:174–89. Doi:
10.1016/j.cobme.2020.02.001
133. Farbod F, Mazloum M. Chapter 5 - Typically used nanomaterials-based noncarbon materials in the
fabrication of biosensors. In: Ensafi AA. Electrochemical Biosensors.Elsevier; 2019. p. 99–133. Doi:
10.1016/B978-0-12-816491-4.00005-X
134. Marx S, Jose MV, Andersen JD, Russell AJ. Electrospun gold nanofiber electrodes for biosensors.Biosens
Bioelectron. 2011;26(6):2981–6. Doi: 10.1016/j.bios.2010.11.050
135. Xu S. Electromechanical biosensors for pathogen detection.Microchim Acta. 2012;178(3):245–60. Doi:
10.1007/s00604-012-0831-4
136. Chen SH, Wu VC, Chuang YC, Lin CS. Using oligonucleotide-functionalized Au nanoparticles to rapidly
detect foodborne pathogens on a piezoelectric biosensor. J Microbiol Methods. 2008;73(1):7–17. Doi:
10.1016/j.mimet.2008.01.004
137. Singh R, Mukherjee MD, Sumana G, Gupta RK, Sood S, Malhotra BD. Biosensors for pathogen detection:
A smart approach towards clinical diagnosis. Sens Actuators B Chem. 2014;197:385–404. Doi:
10.1016/j.snb.2014.03.005
138. Tripathi SM, Bock WJ, Mikulic P, Chinnappan R, Ng A, Tolba M, et al. Long period grating based
biosensor for the detection of Escherichia coli bacteria. Biosens Bioelectron. 2012;35(1):308–12. Doi:
10.1016/j.bios.2012.03.006
139. Chowdhury AD, De A, Chaudhuri CR, Bandyopadhyay K, Sen P. Label free polyaniline based
impedimetric biosensor for detection of E. coli O157:H7 Bacteria. Sens Actuators B Chem. 2012;171–
172:916–23. Doi: 10.1016/j.snb.2012.06.004
140. Wang Y, Alocilja EC. Gold nanoparticle-labeled biosensor for rapid and sensitive detection of bacterial
pathogens. J Biol Eng. 2015;9(1):16. Doi: 10.1186/s13036-015-0014-z
141. Vaisocherová H, Víšová I, Ermini ML, Špringer T, Song XC, Mrázek J, et al. Low-fouling surface plasmon
resonance biosensor for multi-step detection of foodborne bacterial pathogens in complex food samples.
Biosens Bioelectron. 2016;80:84–90. Doi: 10.1016/j.bios.2016.01.040
142. Hernandez R, Valles C, Benito A, Maser W, Rius FX, Riu J. Graphene-based potentiometric biosensor for
the immediate detection of living bacteria. Biosens Bioelectron. 2014;54:553–7. Doi:
10.1016/j.bios.2013.11.053
143. Liu X, Marrakchi M, Xu D, Dong H, Andreescu S. Biosensors based on modularly designed synthetic
peptides for recognition, detection and live/dead differentiation of pathogenic bacteria. Biosens
Bioelectron. 2016;80:9–16. Doi: 10.1016/j.bios.2016.01.041
144. Tak M, Gupta V, Tomar M. Flower-like ZnO nanostructure based electrochemical DNA biosensor for
bacterial meningitis detection. Biosens Bioelectron. 2014;59:200–7. Doi: 10.1016/j.bios.2014.03.036
51
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
52
Sensors 2020, 20, x FOR PEER REVIEW 27 of 32
145. Afsahi S, Lerner MB, Goldstein JM, Lee J, Tang X, Bagarozzi DA, et al. Novel graphene-based biosensor
for early detection of Zika virus infection. Biosens Bioelectron. 2018;100:85–8. Doi:
10.1016/j.bios.2017.08.051
146. Krishna VD, Wu K, Perez AM, Wang JP. Giant Magnetoresistance-based Biosensor for Detection of
Influenza A Virus. Front Microbiol. 2016;7. Doi: 10.3389/fmicb.2016.00400
147. Manzano M, Viezzi S, Mazerat S, Marks RS, Vidic J. Rapid and label-free electrochemical DNA biosensor
for detecting hepatitis A virus. Biosens Bioelectron. 2018;100:89–95. Doi: 10.1016/j.bios.2017.08.043
148. Faria HAM, Zucolotto V. Label-free electrochemical DNA biosensor for zika virus identification. Biosens
Bioelectron. 2019;131:149–55. Doi: 10.1016/j.bios.2019.02.018
149. Singhal C, Khanuja M, Chaudhary N, Pundir CS, Narang J. Detection of chikungunya virus DNA using
two-dimensional MoS 2 nanosheets based disposable biosensor. Sci Rep. 2018;8(1):7734. Doi:
10.1038/s41598-018-25824-8
150. Shakoori Z, Salimian S, Kharrazi S, Adabi M, Saber R. Electrochemical DNA biosensor based on gold
nanorods for detecting hepatitis B virus. Anal Bioanal Chem. 2015;407(2):455–61. Doi: 10.1007/s00216-
014-8303-9
151. Chang YF, Wang WH, Hong YW, Yuan RY, Chen KH, Huang YW, et al. Simple Strategy for Rapid and
Sensitive Detection of Avian Influenza A H7N9 Virus Based on Intensity-Modulated SPR Biosensor and
New Generated Antibody. Anal Chem. 2018;90(3):1861–9. Doi: 10.1021/acs.analchem.7b03934
152. Nuzaihan M, Hashim U, Arshad MK, Kasjoo SR, Rahman SF, Ruslinda AR, et al. Electrical detection of
dengue virus (DENV) DNA oligomer using silicon nanowire biosensor with novel molecular gate control.
Biosens Bioelectron. 2016;83:106–14. Doi: 10.1016/j.bios.2016.04.033
153. Mahato K, Maurya PK, Chandra P. Fundamentals and commercial aspects of nanobiosensors in point-of-
care clinical diagnostics. 3 Biotech. 2018;8(3):149. Doi: 10.1007/s13205-018-1148-8
154. Bahadır EB, Sezgintürk MK. Applications of commercial biosensors in clinical, food, environmental, and
biothreat/biowarfare analyses. Anal Biochem. 2015;478:107–20. Doi: 10.1016/j.ab.2015.03.011
155. Wang J, Chen G, Jiang H, Li Z, Wang X. Advances in nano-scaled biosensors for biomedical applications.
Analyst. 2013;138(16):4427–35. Doi: 10.1039/C3AN00438D
156. Ferreira D, Seca AM, CGA D, Silva AM. Targeting human pathogenic bacteria by siderophores: A
proteomics review. J Proteomics. 2016;145:153–66. Doi: 10.1016/j.jprot.2016.04.006
157. Majdinasab M, Mishra RK, Tang X, Marty JL. Detection of antibiotics in food: New achievements in the
development of biosensors. TrAC Trends Anal Chem. 2020;127:115883. Doi: 10.1016/j.trac.2020.115883
158. Wang YX, Ye ZZ, Si CY, Ying YB. Application of Aptamer Based Biosensors for Detection of Pathogenic
Microorganisms. Chin J Anal Chem. 2012;40(4):634–42. Doi: 10.1016/S1872-2040(11)60542-2
159. Gopal J, Lakshmaiah J, Wu HF. TiO2 nanoparticle assisted mass spectrometry as biosensor of
Staphylococcus aureus, key pathogen in nosocomial infections from air, skin surface and human nasal
passage. Biosens Bioelectron. 2011;27(1):201–6. Doi: 10.1016/j.bios.2011.06.035
160. Rubab M, Shahbaz HM, Olaimat AN, Oh DH. Biosensors for rapid and sensitive detection of
Staphylococcus aureus in food.Biosens Bioelectron. 2018;105:49–57. Doi: 10.1016/j.bios.2018.01.023
161. Suaifan G, Alhogail S, Zourob M. Rapid and low-cost biosensor for the detection of Staphylococcus
aureus. Biosens Bioelectron. 2017;90:230–7. Doi: 10.1016/j.bios.2016.11.047
162. Ahari H, Hedayati M, Akbari B, Kakoolaki S, Hosseini H, Anvar A. Staphylococcus aureus exotoxin
detection using potentiometric nanobiosensor for microbial electrode approach with the effects of pH and
temperature. Int J Food Prop. 2017;20(sup2):1578–87. Doi:10.1080/10942912.2017.1347944
53
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
54
Sensors 2020, 20, x FOR PEER REVIEW 28 of 32
163. Cecchini F, Fajs L, Cosnier S, Marks RS. Vibrio cholerae detection: Traditional assays, novel diagnostic
techniques and biosensors. TrAC Trends Anal Chem. 2016;79:199–209. Doi: 10.1016/j.trac.2016.01.017
164. Narmani A, Kamali M, Amini B, Kooshki H, Amini A, Hasani L. Highly sensitive and accurate detection
of Vibrio cholera O1 OmpW gene by fluorescence DNA biosensor based on gold and magnetic
nanoparticles. Process Biochem. 2018;65:46–54. Doi: 10.1016/j.procbio.2017.10.009
165. Xiao R, Rong Z, Long F, Liu Q. Portable evanescent wave fiber biosensor for highly sensitive detection of
Shigella. Spectrochim Acta A Mol Biomol Spectrosc. 2014;132:1–5. Doi: 10.1016/j.saa.2014.04.090
166. Elahi N, Baghersad MH, Kamali M. Precise, direct, and rapid detection of Shigella Spa gene by a novel
unmodified AuNPs-based optical genosensing system.J Microbiol Methods. 2019;162:42–9. Doi:
10.1016/j.mimet.2019.05.007
167. Dao TN, Yoon J, Jin CE, Koo B, Han K, Shin Y, et al. Rapid and sensitive detection of Salmonella based on
microfluidic enrichment with a label-free nanobiosensing platform. Sens Actuators B Chem.
2018;262:588–94. Doi: 10.1016/j.snb.2017.12.190
168. Zhang X, Wu D, Zhou X, Yu Y, Liu J, Hu N, et al. Recent progress in the construction of nanozyme-based
biosensors and their applications to food safety assay. TrAC Trends Anal Chem. 2019;121:115668. Doi:
10.1016/j.trac.2019.115668
169. Barroso TG, Martins RC, Fernandes E, Cardoso S, Rivas J, Freitas PP. Detection of BCG bacteria using a
magnetoresistive biosensor: A step towards a fully electronic platform for tuberculosis point-of-care
detection. Biosens Bioelectron. 2018;100:259–65. Doi: 10.1016/j.bios.2017.09.004
170. Mandal HS, Su Z, Ward A, Tang X. Carbon Nanotube Thin Film Biosensors for Sensitive and
Reproducible Whole Virus Detection. Theranostics. 2012;2(3):251–7. Doi:10.7150/thno.3726
171. Banga I, Tyagi R, Shahdeo D, Gandhi S. Chapter 1 - Biosensors and Their Application for the Detection of
Avian Influenza Virus. In: Maurya PK, Singh S. Nanotechnology in Modern Animal Biotechnology.
Elsevier; 2019.p. 1–16. Doi: 10.1016/B978-0-12-818823-1.00001-6
172. Nidzworski D, Pranszke P, Grudniewska M, Król E, Gromadzka B. Universal biosensor for detection of
influenza virus. Biosens Bioelectron. 2014;59:239–42. Doi: 10.1016/j.bios.2014.03.050
173. Tepeli Y, Ülkü A. Electrochemical biosensors for influenza virus a detection: The potential of adaptation
of these devices to POC systems. Sens Actuators B Chem. 2018;254:377–84. Doi: 10.1016/j.snb.2017.07.126
174. Hassanpour S, Baradaran B, Hejazi M, Hasanzadeh M, Mokhtarzadeh A, de la Guardia M. Recent trends
in rapid detection of influenza infections by bio and nanobiosensor. TrAC Trends Anal Chem.
2018;98:201–15. Doi: 10.1016/j.trac.2017.11.012
175. Kaushik A, Tiwari S, Jayant RD, Vashist A, Nikkhah R, El-Hage N, et al. Electrochemical Biosensors for
Early Stage Zika Diagnostics. Trends Biotechnol. 2017;35(4):308–17. Doi: 10.1016/j.tibtech.2016.10.001
176. Jiang L, Luo J, Dong W, Wang C, Jin W, Xia Y, et al. Development and evaluation of a polydiacetylene
based biosensor for the detection of H5 influenza virus. J Virol Methods. 2015;219:38–45. Doi:
10.1016/j.jviromet.2015.03.013
177. Rojas M, Monsalve DM, Pacheco Y, Acosta Y, Ramírez C, Ansari AA, et al. Ebola virus disease: An
emerging and re-emerging viral threat. J Autoimmun. 2020;106:102375. Doi. 10.1016/j.jaut.2019.102375
178. Kharsany AB, McKinnon LR, Lewis L, Cawood C, Khanyile D, Maseko DV, et al. Population prevalence
of sexually transmitted infections in a high HIV burden district in KwaZulu-Natal, South Africa:
Implications for HIV epidemic control. Int J Infect Dis. 2020;98:130–7. Doi: 10.1016/j.ijid.2020.06.046
179. Mittler E, Dieterle ME, Kleinfelter LM, Slough MM, Chandran K, Jangra RK. Chapter Six - Hantavirus
entry: Perspectives and recent advances. In: Kielian M, Mettenleiter TC, Roossinck MJ. Advances in Virus
Research.Academic Press; 2019. p. 185–224. Doi: 10.1016/bs.aivir.2019.07.002
55
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
56
Sensors 2020, 20, x FOR PEER REVIEW 29 of 32
180. Nicastri E, Kobinger G, Vairo F, Montaldo C, Mboera LE, Ansunama R, et al. Ebola Virus Disease:
Epidemiology, Clinical Features, Management, and Prevention. Infect Dis Clin North Am. 2019;33(4):953–
76. Doi: 10.1016/j.idc.2019.08.005
181. Mérens A, Bigaillon C, Delaune D. Ebola virus disease: Biological and diagnostic evolution from 2014 to
2017. Médecine Mal Infect. 2018;48(2):83–94. Doi: 10.1016/j.medmal.2017.11.002
182. Murphy CN. Recent Advances in the Diagnosis and Management of Ebola Virus Disease. Clin Microbiol
Newsl. 2019;41(21):185–9. Doi: 10.1016/j.clinmicnews.2019.10.002
183. Walldorf JA, Cloessner EA, Hyde TB, MacNeil A, Bennett SD, Carter RJ, et al. Considerations for use of
Ebola vaccine during an emergency response. Vaccine. 2019 Nov;37(48):7190–200. Doi:
10.1016/j.vaccine.2017.08.058
184. Kaushik A, Tiwari S, Dev Jayant R, Marty A, Nair M. Towards detection and diagnosis of Ebola virus
disease at point-of-care. Biosens Bioelectron. 2016;75:254–72. Doi: 10.1016/j.bios.2015.08.040
185. Ilkhani H, Farhad S. A novel electrochemical DNA biosensor for Ebola virus detection.Anal Biochem.
2018;557:151–5. Doi: 10.1016/j.ab.2018.06.010
186. Baca JT, Severns V, Lovato D, Branch DW, Larson RS. Rapid Detection of Ebola Virus with a Reagent-
Free, Point-of-Care Biosensor. Sensors. 2015;15(4):8605–14. Doi: 10.3390/s150408605
187. Iordache L, Launay O, Bouchaud O, Jeantils V, Goujard C, Boue F, et al. Autoimmune diseases in HIV-
infected patients: 52 cases and literature review. Autoimmun Rev. 2014;13(8):850–7. Doi:
10.1016/j.autrev.2014.04.005
188. Nandi S, Mondal A, Roberts A, Gandhi S. Chapter One - Biosensor platforms for rapid HIV detection. In:
Makowski GS. Advances in Clinical Chemistry.Elsevier; 2020. p. 1–34. Doi: 10.1016/bs.acc.2020.02.001
189. Tavakoli A, Karbalaie MH, Keshavarz M, Ghaffari H, Asoodeh A, Monavari SH, et al. Current diagnostic
methods for HIV. Future Virol. 2017;12(3):141–55. Doi: 10.2217/fvl-2016-0096
190. Shafiee H, Lidstone EA, Jahangir M, Inci F, Hanhauser E, Henrich TJ, et al. Nanostructured Optical
Photonic Crystal Biosensor for HIV Viral Load Measurement. Sci Rep. 2014;4(1):4116. Doi:
10.1038/srep04116
191. Gong Q, Han H, Yang H, Zhang M, Sun X, Liang Y, et al. Sensitive electrochemical DNA sensor for the
detection of HIV based on a polyaniline/graphene nanocomposite. J Materiomics. 2019;5(2):313–9. Doi:
10.1016/j.jmat.2019.03.004
192. Vetcha S, Wilkins E, Yates T, Hjelle B. Rapid and sensitive handheld biosensor for detection of
hantavirus antibodies in wild mouse blood samples under field conditions. Talanta. 2002;58(3):517–28.
Doi: 10.1016/S0039-9140(02)00307-7
193. Jiang H, Zheng X, Wang L, Du H, Wang P, Bai X. Hantavirus infection: a global zoonotic challenge. Virol
Sin. 2017;32(1):32–43. Doi:10.1007/s12250-016-3899-x
194. Gogola JL, Martins G, Caetano FR, Ricciardi T, Duarte CN, Marcolino LH, et al. Label-free
electrochemical immunosensor for quick detection of anti-hantavirus antibody. J Electroanal Chem.
2019;842:140–5. Doi: 10.1016/j.jelechem.2019.04.066
195. Martins G, Gogola JL, Caetano FR, Kalinke C, Jorge TR, Duarte CN, et al. Quick electrochemical
immunoassay for hantavirus detection based on biochar platform. Talanta. 2019;204:163–71. Doi:
10.1016/j.talanta.2019.05.101
196. Farzin L, Shamsipur M, Samandari L, Sheibani S. HIV biosensors for early diagnosis of infection: The
intertwine of nanotechnology with sensing strategies. Talanta. 2020;206:120201. Doi:
10.1016/j.talanta.2019.120201
57
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
58
Sensors 2020, 20, x FOR PEER REVIEW 30 of 32
197. Panghal A, Flora SJ. Chapter 4 - Viral agents including threat from emerging viral infections. In: Flora SJ,
Pachauri V. Handbook on Biological Warfare Preparedness. Academic Press; 2020. p. 65–81. Doi:
10.1016/B978-0-12-812026-2.00004-9
198. Al-Rohaimi AH, Al-Otaibi F. Novel SARS-CoV-2 outbreak and COVID19 disease; a systemic review on
the global pandemic. Genes Dis. 2020. Doi: 10.1016/j.gendis.2020.06.004
199. Khan MZ, Hasan MR, Hossain SI, Ahommed MS, Daizy M. Ultrasensitive detection of pathogenic
viruses with electrochemical biosensor: State of the art. Biosens Bioelectron. 2020;112431. Doi:
10.1016/j.bios.2020.112431
200. Ji T, Liu Z, Wang G, Guo X, Akbar khan S, Lai C, et al. Detection of COVID-19: A review of the current
literature and future perspectives. Biosens Bioelectron. 2020;112455. Doi: 10.1016/j.bios.2020.112455
201. Sheikhzadeh E, Eissa S, Ismail A, Zourob M. Diagnostic techniques for COVID-19 and new
developments. Talanta. 2020;220:121392. Doi: 10.1016/j.talanta.2020.121392
202. Yuan X, Yang C, He Q, Chen J, Yu D, Li J, et al. Current and Perspective Diagnostic Techniques for
COVID-19. ACS Infectious Diseases. 2020;6(8):1998-2006. Doi: 10.1021/acsinfecdis.0c00365
203. Jalandra R, Yadav AK, Verma D, Dalal N, Sharma M, Singh R, et al. Strategies and perspectives to
develop SARS-CoV-2 detection methods and diagnostics.Biomed Pharmacother. 2020;129:110446. Doi:
10.1016/j.biopha.2020.110446
204. Ward S, Lindsley A, Courter J, Assa’ad A. Clinical testing for COVID-19. J Allergy Clin Immunol.
2020;146(1):23–34. Doi: 10.1016/j.jaci.2020.05.012
205. Kumar R, Nagpal S, Kaushik S, Mendiratta S. COVID-19 diagnostic approaches: different roads to the
same destination. VirusDisease. 2020;31(2):97–105. Doi: 10.1007/s13337-020-00599-7
206. Santiago I. Trends and Innovations in Biosensors for COVID 19 Mass Testing. ChemBioChem. 2020;21:1-
11. Doi: 10.1002/cbic.202000250
207. Kaur M, Tiwari S, Jain R. Protein based biomarkers for non-invasive Covid-19 detection. Sens Bio-Sens
Res. 2020;29:100362. Doi: 10.1016/j.sbsr.2020.100362
208. Choi JR. Development of Point-of-Care Biosensors for COVID-19. Front Chem. 2020;8:517. Doi:
10.3389/fchem.2020.00517
209. Sheridan C. Fast, portable tests come online to curb coronavirus pandemic. Nat Biotechnol.
2020;38(5):515–8. Doi:10.1038/d41587-020-00010-2
210. Zhifeng J, Feng A, Li T. Consistency analysis of COVID-19 nucleic acid tests and the changes of lung CT.
J Clin Virol. 2020;127:104359. Doi: 10.1016/j.jcv.2020.104359
211. Li Z, Yi Y, Luo X, Xiong N, Liu Y, Li S, et al. Development and clinical application of a rapid IgM IgG
combined antibody test for SARS CoV 2 infection diagnosis. J Med Virol. 2020;2(9):1518-24. Doi:‐ ‐
10.1002/jmv.25727
212. Thevarajan I, Nguyen T, Koutsakos M, Druce J, Caly L, van de Sandt C, et al. Breadth of concomitant
immune responses prior to patient recovery: a case report of non-severe COVID-19. Nature Medicine.
2020;26:453-455. Doi: 10.1038/s41591-020-0819-2
213. Du Z, Zhu F, Guo F, Yang B, Wang T. Detection of antibodies against SARS-CoV-2 in patients with
COVID-19. J Med Virol. 2020. Doi: 10.1002/jmv.25820
214. Health C for D and R. EUA Authorized Serology Test Performance. FDA. 2020. Available from:
https://www.fda.gov/medical-devices/coronavirus-disease-2019-covid-19-emergency-use-
authorizations-medical-devices/eua-authorized-serology-test-performance
215. Ghaffari A, Meurant R, Ardakani A. COVID-19 Serological Tests: How Well Do They Actually Perform?
Diagnostics. 2020;10(7):453. Doi: 10.3390/diagnostics10070453
59
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
60
Sensors 2020, 20, x FOR PEER REVIEW 31 of 32
216. Morales E, Dincer C. The impact of biosensing in a pandemic outbreak: COVID-19. Biosens Bioelectron.
2020;163:112274. Doi: 10.1016/j.bios.2020.112274
217. Yin J, Zou Z, Hu Z, Zhang S, Zhang F, Wang B, et al. A “sample-in-multiplex-digital-answer-out” chip
for fast detection of pathogens. Lab Chip. 2020;20(5):979–86. Doi: 10.1039/C9LC01143A
218. Cady NC, Fusco V, Maruccio G, Primiceri E, Batt CA. 12 - Micro- and nanotechnology-based approaches
to detect pathogenic agents in food. In: Grumezescu AM. Nanobiosensors.Academic Press; 2017. p. 475–
510. Doi: 10.1016/B978-0-12-804301-1.00012-6
219. Zhuang J, Yin J, Lv S, Wang B, Mu Y. Advanced “lab-on-a-chip” to detect viruses – Current challenges
and future perspectives. Biosens Bioelectron. 2020;163:112291. Doi: 10.1016/j.bios.2020.112291
220. Mavrikou S, Moschopoulou G, Tsekouras V, Kintzios S. Development of a Portable, Ultra-Rapid and
Ultra-Sensitive Cell-Based Biosensor for the Direct Detection of the SARS-CoV-2 S1 Spike Protein
Antigen. Sensors. 2020;20(11):3121. Doi: 10.3390/s20113121
221. Bhalla N, Pan Y, Yang Z, Farokh A. Opportunities and Challenges for Biosensors and Nanoscale
Analytical Tools for Pandemics: COVID-19. ACS Nano. 2020;14(7):7783-7807. Doi:
10.1021/acsnano.0c04421
222. Palestino G, Garcia I, Gonzalez O, Rosales S. Can nanotechnology help in the fight against COVID-19?
Exp Rev Anti-Innefect Ther. 2020. Doi: 10.1080/14787210.2020.1776115
223. Vermisoglou E, Panáček D, Jayaramulu K, Pykal M, Frébort I, Kolář M, et al. Human virus detection with
graphene-based materials. Biosens Bioelectron. 2020;112436. Doi: 10.1016/j.bios.2020.112436
224. Palmieri V, Papi M. Can graphene take part in the fight against COVID-19? Nano Today. 2020;33:100883.
Doi: 10.1016/j.nantod.2020.100883
225. Seo G, Lee G, Kim MJ, Baek SH, Choi M, Ku KB, et al. Rapid Detection of COVID-19 Causative Virus
(SARS-CoV-2) in Human Nasopharyngeal Swab Specimens Using Field-Effect Transistor-Based
Biosensor. ACS Nano. 2020;14(4):5135–42. Doi: 10.1021/acsnano.0c02823
226. Cui F, Zhou HS. Diagnostic methods and potential portable biosensors for coronavirus disease 2019.
Biosens Bioelectron. 2020;165:112349. Doi: 10.1016/j.bios.2020.112349
227. Murugan D, Bhatia H, Sai VV, Satija J. P-FAB: A Fiber-Optic Biosensor Device for Rapid Detection of
COVID-19. Trans Indian Natl Acad Eng. 2020;5:211-215. Doi: 10.1007/s41403-020-00122-w
228. Zhu X, Wang X, Han L, Chen T, Wang L, Li H, et al. Multiplex reverse transcription loop-mediated
isothermal amplification combined with nanoparticle-based lateral flow biosensor for the diagnosis of
COVID-19. Biosens Bioelectron. 2020;166:112437. Doi: 10.1016/j.bios.2020.112437
229. Qiu G, Gai Z, Tao Y, Schmitt J, Kullak GA, Wang J. Dual-Functional Plasmonic Photothermal Biosensors
for Highly Accurate Severe Acute Respiratory Syndrome Coronavirus 2 Detection. ACS Nano.
2020;14(5):5268–77. Doi: 10.1021/acsnano.0c02439
230. Mujawar MA, Gohel H, Bhardwaj SK, Srinivasan S, Hickman N, Kaushik A. Nano-enabled biosensing
systems for intelligent healthcare: towards COVID-19 management. Mater Today Chem. 2020;17:100306.
Doi: 10.1016/j.mtchem.2020.100306
231. Qin Z, Peng R, Baravik IK, Liu X. Fighting COVID-19: Integrated Micro- and Nanosystems for Viral
Infection Diagnostics. Matter. 2020. Doi: 10.1016/j.matt.2020.06.015
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.
61
1116
1117
1118
1119
1120
1121
1122
1123
1124
1125
1126
1127
1128
1129
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
62
Sensors 2020, 20, x FOR PEER REVIEW 32 of 32
© 2020 by the authors. Submitted for possible open access publication under the terms and
conditions of the Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/).
63
1158
64
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
A nanocomposite of polyaniline/graphene (PAN/GN) was prepared using reverse-phase polymerization. The nanocomposite material was dropcast onto a glassy carbon electrode (GCE). Then, a single-stranded DNA (ssDNA) probe for HIV-1 gene detection was immobilized on the modified electrode, and the negative charged phosphate backbone of the HIV-1 was bound to the modified electrode surface via π-π∗ stacking interactions. The hybridization between the ssDNA probe and the target HIV-1 formed double-stranded DNA (dsDNA), and the electron transfer resistance of the electrode was measured using impedimetric studies with a [Fe(CN)6]3-/4- redox couple. Under the optimized experimental conditions, the change of the impedance value was linearly related to the logarithm of the concentration of HIV genes in the range from 5.0 × 10⁻¹⁶ M to 1.0 × 10⁻¹⁰ M (R = 0.9930), and the HIV sensor exhibited a lower detection limit of 1.0 × 10⁻¹⁶ M (S/N = 3). The results show that this biosensor presented wonderful selectivity, sensitivity and specificity for HIV-1 gene detection. Thus, this biosensor provides a new method for the detection of HIV gene fragments.
Article
Full-text available
Since late December 2019, the coronavirus pandemic (COVID-19; previously known as 2019-nCoV) caused by the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has been surging rapidly around the world. With more than 13,000,000 confirmed cases, the world faces an unprecedented economic, social, and healthy impact. Early, rapid, sensitive, and accurate diagnosis of viral infection provides rapid responses for public health surveillance, prevention, and control of contagious diffusion. More than 30% of the confirmed cases are asymptomatic and the high false-negative rate (FNR) of a single assay requires the development of novel diagnostic techniques, combinative approaches, sampling from different locations, and consecutive detection. The recurrence of discharged patients indicates the need for long-term monitoring and tracking. Diagnostic and therapeutic methods are evolving with a deeper understanding of virus pathology and the potential for relapse. In this review, a comprehensive summary and comparison of different SARS-CoV-2 diagnostic methods are provided for researchers and clinicians to develop appropriate strategies for the timely and effective detection of SARS-CoV-2. The survey of current biosensors and diagnostic devices for viral nucleic acid, protein, particle, and chest tomography will provide insight into the development of novel perspective techniques for the diagnosis of COVID-19.
Article
Full-text available
In only a few months after initial discovery in Wuhan, China, SARS-CoV-2 and the associated coronavirus disease 2019 (COVID-19) have become a global pandemic causing significant mortality and morbidity and implementation of strict isolation measures. In the absence of vaccines and effective therapeutics, reliable serological testing must be a key element of public health policy to control further spread of the disease and gradually remove quarantine measures. Serological diagnostic tests are being increasingly used to provide a broader understanding of COVID-19 incidence and to assess immunity status in the population. However, there are discrepancies between claimed and actual performance data for serological diagnostic tests on the market. In this study, we conducted a review of independent studies evaluating the performance of SARS-CoV-2 serological tests. We found significant variability in the accuracy of marketed tests and highlight several lab-based and point-of-care rapid serological tests with high levels of performance. The findings of this review highlight the need for ongoing independent evaluations of commercialized COVID-19 diagnostic tests.
Article
Soft and stretchable biosensors have recently gained considerable attention owing to their unique capability of seamless integration with the intrinsically soft and curvilinear biological tissues for wearable/implantable health monitoring remotely. In this review, we aim to focus on surveying the recent achievements in monitoring key biological health vital markers including glucose, lactate, and cationic ions by electrochemical means. First, we discuss the soft and stretchable electrochemical electrodes (namely working, counter, and reference electrodes) from the perspectives of materials and design strategies. Then, we cover viable detection methods and modalities used in soft and stretchable biosensing systems. This is followed by description of the exciting applications of soft electrochemical biosensing systems in non-intrusive personal healthcare monitoring and in real-time monitoring of cells and tissues. This review concludes by the discussion of the challenges, opportunities, and future perspectives of soft and stretchable electrochemical biosensors.
Article
Last few decades, viruses are a real menace to human safety. Therefore, the rapid identification of viruses should be one of the best ways to prevent an outbreak and important implications for medical healthcare. The recent outbreak of coronavirus disease (COVID-19) is an infectious disease caused by a newly discovered coronavirus which belongs to the single-stranded, positive-strand RNA viruses. The pandemic dimension spread of COVID-19 poses a severe threat to the health and lives of seven billion people worldwide. There is a growing urgency worldwide to establish a point-of-care device for the rapid detection of COVID-19 to prevent subsequent secondary spread. Therefore, the need for sensitive, selective, and rapid diagnostic devices plays a vital role in selecting appropriate treatments and to prevent the epidemics. During the last decade, electrochemical biosensors have emerged as reliable analytical devices and represent a new promising tool for the detection of different pathogenic viruses. This review summarizes the state of the art of different virus detection with currently available electrochemical detection methods. Moreover, this review discusses different fabrication techniques, detection principles, and applications of various virus biosensors. Future research also looks at the use of electrochemical biosensors regarding a potential detection kit for the rapid identification of the COVID-19.
Article
Our recent experience of the COVID-19 pandemic has highlighted the importance of easy-to-use, quick, cheap, sensitive and selective detection of virus pathogens for the efficient monitoring and treatment of virus diseases. Early detection of viruses provides essential information about possible efficient and targeted treatments, prolongs the therapeutic window and hence reduces morbidity. Graphene is a lightweight, chemically stable and conductive material that can be successfully utilized for the detection of various virus strains. The sensitivity and selectivity of graphene can be enhanced by its functionalization or combination with other materials. Introducing suitable functional groups and/or counterparts in the hybrid structure enables tuning of the optical and electrical properties, which is particularly attractive for rapid and easy-to-use virus detection. In this review, we cover all the different types of graphene-based sensors available for virus detection, including, e.g., photoluminescence and colorimetric sensors, and surface plasmon resonance biosensors. Various strategies of electrochemical detection of viruses based on, e.g., DNA hybridization or antigen-antibody interactions, are also discussed. We summarize the current state-of-the-art applications of graphene-based systems for sensing a variety of viruses, e.g., SARS-CoV-2, influenza, dengue fever, hepatitis C virus, HIV, rotavirus and Zika virus. General principles, mechanisms of action, advantages and drawbacks are presented to provide useful information for the further development and construction of advanced virus biosensors. We highlight that the unique and tunable physicochemical properties of graphene-based nanomaterials make them ideal candidates for engineering and miniaturization of biosensors.
Article
The rapid spread of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has led to the coronavirus disease 2019 (COVID-19) worldwide pandemic. This unprecedented situation has garnered worldwide attention. An effective strategy for controlling the COVID-19 pandemic is to develop highly accurate methods for the rapid identification and isolation of SARS-CoV-2 infected patients. Many companies and institutes are therefore striving to develop effective methods for the rapid detection of SARS-CoV-2 ribonucleic acid (RNA), antibodies, antigens, and the virus. In this review, we summarize the structure of the SARS-CoV-2 virus, its genome and gene expression characteristics, and the current progression of SARS-CoV-2 RNA, antibodies, antigens, and virus detection. Further,we discuss the reasons for the observed false-negative and false-positive RNA and antibody detection results in practical clinical applications. Finally, we provide a review of the biosensors which hold promising potential for point-of-care detection of COVID-19 patients. This review thereby provides general guidelines for both scientists in the biosensing research community and for those in the biosensor industry to develop a highly sensitive and accurate point-of-care COVID-19 detection system, which would be of enormous benefit for controlling the current COVID-19 pandemic.
Article
COVID-19 pandemic is a serious global health issue today due to the rapid human to human transmission of SARS-CoV-2, a new type of coronavirus that causes fatal pneumonia. SARS -CoV-2 has a faster rate of transmission than other coronaviruses such as SARS and MERS and until now there are no approved specific drugs or vaccines for treatment. Thus, early diagnosis is crucial to prevent the extensive spread of the disease. The reverse transcription-polymerase chain reaction (RT-PCR) is the most routinely used method until now to detect SARS-CoV-2 infections. However, several other faster and accurate assays are being developed for the diagnosis of COVID-19 aiming to control the spread of infection through the identification of patients and immediate isolation. In this review, we will discuss the various detection methods of the SARS-CoV-2 virus including the recent developments in immunological assays, amplification techniques as well as biosensors.
Article
There is a growing demand to protect food products against the hazard of microbes and their toxins. To satisfy such goals, it is important to develop highly sensitive, reliable, sophisticated, rapid, and cost-effective sensing techniques such as electrochemical sensors/biosensors. Although diverse forms of nanomaterials (NMs)-based electrochemical sensing methods have been introduced in markets, the reliability of commercial products is yet insufficient to meet the practical goal. In this review, we focused on: 1) sources of pathogenic microbes and their toxins; 2) possible routes of their entrainment in food, and 3) current development of NM-based biosensors to realize real-time detection of the target analytes. At last, future prospects and challenges in this research field are discussed.
Article
Electrospinning is a novel and sophisticated technique for the production of nanofibers with high surface area, extreme porous structure, small pore size, and surface morphologies that make them suitable for biomedical and bioengineering applications, which can provide solutions to current drug delivery issues of poorly water-soluble drugs. Electrospun nanofibers can be obtained through different methods asides from the conventional one, such as coaxial, multi-jet, side by side, emulsion, and melt electrospinning. In general, the application of an electric potential to a polymer solution causes a charged liquid jet that moves downfield to an oppositely charged collector, where the nanofibers are deposited. Plenty of polymers that differ in their origin, degradation character and water affinity are used during the process. Physicochemical properties of the drug, polymer(s), and solvent systems need to be addressed to guarantee successful manufacturing. Therefore, this review summarizes the recent progress in electrospun nanofibers for their use as a nanotechnological tool for dissolution optimization and drug delivery systems for poorly water-soluble drugs.