DataPDF Available
Journal Pre-proofs
Make it clean, make it safe: A review on virus elimination via adsorption
Lotfi Sellaoui, Michael Badawi, Antonio Monari, Tetiana Tatarchuk, Sonia
Jemli, Guilherme Luiz Dotto, Adrian Bonilla-Petriciolet, Zhuqi Chen
PII: S1385-8947(21)00280-1
DOI: https://doi.org/10.1016/j.cej.2021.128682
Reference: CEJ 128682
To appear in: Chemical Engineering Journal
Received Date: 14 September 2020
Revised Date: 21 December 2020
Accepted Date: 13 January 2021
Please cite this article as: L. Sellaoui, M. Badawi, A. Monari, T. Tatarchuk, S. Jemli, G. Luiz Dotto, A. Bonilla-
Petriciolet, Z. Chen, Make it clean, make it safe: A review on virus elimination via adsorption, Chemical
Engineering Journal (2021), doi: https://doi.org/10.1016/j.cej.2021.128682
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2021 Elsevier B.V. All rights reserved.
1
Make it clean, make it safe: A review on virus elimination via adsorption
Lotfi Sellaoui 1,2*, Michael Badawi 3**, Antonio Monari 3, Tetiana Tatarchuk 4, Sonia Jemli 5,6,
Guilherme Luiz Dotto 7, Adrian Bonilla-Petriciolet 8, Zhuqi Chen1,2***
1 Department of Environmental Engineering, School of Environmental Science and Engineering,
Huazhong University of Science and Technology, Wuhan 430074, PR China
2 Key Laboratory of Material Chemistry for Energy Conversion and Storage, Ministry of
Education, Hubei Key Laboratory of Material Chemistry and Service Failure, School
of Chemistry and Chemical Engineering, Huazhong University of Science and Technology,
Wuhan 430074, PR China
3 Laboratoire de Physique et Chimie Théoriques LPCT UMR CNRS 7019, Université de
Lorraine, Vandœuvre-lès-Nancy, France
4 Educational and Scientific Center of Material Science and Nanotechnology, Vasyl Stefanyk
Precarpathian National University, Ivano-Frankivsk, 76018, Ukraine
5 Laboratory of Microbial Biotechnology, Enzymatic and Biomolecules (LMBEB), Centre of
Biotechnology of Sfax, University of Sfax, Tunisia
6 Faculty of Sciences of Sfax, Biology department, University of Sfax, Tunisia.
7 Chemical Engineering Department, Federal University of Santa Maria–UFSM, 1000, Roraima
Avenue, 97105-900 Santa Maria, RS, Brazil
8 Instituto Tecnológico de Aguascalientes, Aguascalientes, 20256, México
Corresponding authors:
*Lotfi Sellaoui, sellaouilotfi@yahoo.fr , **Michael Badawi, michael.badawi@univ-lorraine.fr
***Zhuqi Chen, zqchen@hust.edu.cn
2
Abstract
Recently, the potential dangers of viral infection transmission through water and air have
become the focus of worldwide attention, via the spread of COVID-19 pandemic. The
occurrence of large-scale outbreaks of dangerous infections caused by unknown pathogens and
the isolation of new pandemic strains require the development of improved methods of viruses’
inactivation. Viruses are not stable self-sustaining living organisms and are rapidly inactivated
on isolated surfaces. However, water resources and air can participate in the pathogens’
diffusion, stabilization, and transmission. Viruses inactivation and elimination by adsorption are
relevant since they can represent an effective and low-cost method to treat fluids, and hence limit
the spread of pathogen agents. This review analyzed the interaction between viruses and carbon-
based, oxide-based, porous materials and biological materials (e.g., sulfated polysaccharides and
cyclodextrins). It will be shown that these adsorbents can play a relevant role in the viruses
removal where water and air purification mostly occurring via electrostatic interactions.
However, a clear systematic vision of the correlation between the surface potential and the
adsorption capacity of the different filters is still lacking and should be provided to achieve a
better comprehension of the global phenomenon. The rationalization of the adsorption capacity
may be achieved through a proper physico-chemical characterization of new adsorbents,
including molecular modeling and simulations, also considering the adsorption of virus-like
particles on their surface. As a most timely perspective, the results on this review present
potential solutions to investigate coronaviruses and specifically SARS-CoV-2, responsible of the
COVID-19 pandemic, whose spread can be limited by the efficient disinfection and purification
of closed-spaces air and urban waters.
Keywords: coronavirus; SARS; wastewater; porous; carbon; MOF
3
1. Introduction
Viruses are ubiquitous in nature and consequently their interactions with superior organism and
human beings are constant. Overall, the genetic diversity of viruses, their capabilities to change
and adapt, as well as their significant presence in nature are amazing. It has been estimated that
the total number of viral particles is significantly higher than the number of all cells of all the
organisms on Earth combined [1]. A virus is defined as a non-cellular infectious agent, and the
fundamental question whether they may or not be classified as living organisms is still debated.
Today, only 6,000 viruses are known while several millions viral strains exist but are unknown
[2]. The extreme diversity of the viral populations is also related to the problems encountered in
viruses’ eradication and to the various public health threats. Indeed, viruses are particularly
difficult to eliminate because they are continuous evolving under natural selection pressure,
allowing them to counteract the protection offered by vaccines or specifically targeted drugs.
Viral particles cannot subsist independently, and the first step of the diffusion is the infection of
a cell, that is invaded with the viral genomic and proteic material.
Almost all known viruses have their own specific target in a living organism, which implies the
existence of a specific receptor on the cell surface where the virus attaches itself to gain entry to
the cell [2], the preferred viral receptors also determine the cell types that will most probably
suffer from the infection. Once inside the cell the viral genome is replicated and translated, also
exploiting the cellular machinery, resulting in the production and maturation of novel viral
particles that can be expelled from the cell, usually killing the latter in the process, and
contribute to the further diffusion of the agent. Viral transmission is mostly achieved by the
expulsion of viral particle with bodily fluid or respiration and the subsequent contact with other
organisms. Normally, and once again due to the receptor specificity, the viral transmission is
mostly happening between organisms of the same, or closely related, species. However, the virus
mutation capability can also lead to zoonosis, i.e. overcoming the interspecies barrier greatly
4
facilitating the transmissibility and contagious capacity. Indeed, zoonosis is usually recognized
as a crucial step in the early development of uncontrolled epidemics and outbreaks.
It is nowadays well established that viruses may represent a serious threat to public health, since
they can be extremely pathogenic and be at the origin of extremely dangerous diseases [1,3,4].
Indeed, since ancient times mankind had to cope with the diffusion of epidemic and pandemic,
whose periodic outbreaks have marked many historical periods. The constant threat of viral-
based diseases is also witnessed by the fact that in many ancient cultures its spreading was
attributed to the angers of Gods, see for instance Sekhmet in ancient Egypt or Artemis in
classical Greece. Viral-based diseases may indeed severely affect different organs, and their
spread is favored by the possibility of human-to-human contagion, that is determined by the
inherent viral infectivity rate.
Every year, humanity is facing challenges with new types of viruses that threaten the human
health and can cause epidemics and, in the worse scenario, pandemics [2], the most common,
and relatively innocent one being seasonal flu. A more dangerous example is constituted by the
Middle East respiratory coronavirus syndrome (MERS coronavirus) appeared in the spring of
2015 in South Korea. The outbreak took the South Korean authorities by surprise and forced
them to take urgent epidemiological measures [2]. MERS mortality ratio amounts to more than
35 % and, as stated in the World Health Organization (WHO) Newsletter, “there is currently no
specific treatment or vaccine for this disease”. This example alone justify the extreme
importance and interest devoted to the research on viruses and its vital importance to maintain
healthy environments for the society [5].
In 2020, humanity is facing new global social and economic challenges induced by the
apparition of a novel coronavirus in late 2019. The first cases of an unknown coronavirus were
recorded on 31th December in the Chinese city of Wuhan, which has a population of nearly 12
million [6]. The effects due to this disease were rapidly observed in mainland China neighboring
5
countries: Thailand, Japan, South Korea and Taiwan [7]. Since then, the virus dissemination has
continued, with the epicenter moving from Asia, Europe and later Americas, the new
coronavirus was later named severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2) to
differentiate it from the original SARS-CoV appeared in 2003, while the associated disease has
been styled as COVID-19. On March 11th 2020, WHO declared COVID-19 a pandemic [7]. As
late as November 18th 2020, more than 55 million people have been infected worldwide, new
daily cases and deaths are still globally increasing, although the original Asian hot-spots seem
under relative control. COVID-19 affects the respiratory system, the gastrointestinal tract or the
nervous system; it can cause bronchitis, pneumonia and, in severe cases, death [1,5,6]. SARS-
CoV-2 originally developed in animals, probably pangolins, which acted as the initial main
carriers. However, through zoonosis, the transmissibility via human-to-human contact became
possible and represent today the main contagious route [4], while recent studies have also
pointed to the possible airborne transmission of this virus via aerosols [8]. The mortality ratio of
SARS-CoV-2 is relatively low, especially when compared with SARS-CoV or MERS, and is
estimated to not exceed 1-2 %, severe complications and death usually appearing in aged patients
or in subject presenting significant comorbidity, such as diabetes or hypertension. However, the
transmissibility and diffusivity of the virus are extremely high, and are also facilitated by the
presence of asymptomatic, or barely symptomatic carriers contributing to its spreading.
COVID-19 pandemic and the effects to the related containing measures on the different aspects
of the daily life have been clearly highlighted, also towards the general public, the importance of
developing new strategies to minimize the corresponding infection risks via the reduction of
exposure routes for this and other viruses and to generate healthy environments. The
quantification of coronavirus particles in wastewater is indeed an interesting and original
approach to the study of the spread of the pathogens [9], since SARS-CoV-2 particles have been
found in wastewater in a quantity correlating well with the epidemic growing ratio [10], pushing
6
many countries to begin a more systematic wastewater monitoring [11]. Indeed, the increases of
the number of SARS-CoV-2 particles in wastewater occur even before COVID-19 symptoms
appear in patients, hence its monitoring may allow the early detection of possible epidemic
outbreaks [12]. Even if interesting such an approach requires the use of concentration method to
overcome the low concentration of viral material and it relies on the use of adsorbents such as
anion exchange resin [13].
It is known that the dissemination of viruses can take place by different routes including direct
contact with infected animals and humans, or through dispersion of viral particles in water and
air [3,14–16]. As a consequence, it is necessary to establish high-quality water and air treatment
to contribute to the reduction of virus dissemination, lowering the infectivity and transmissibility
ratio, and hence preventing serious epidemic outbreaks [17,18]. Thus, to be effective and to
respect the public-health standards, purification of water and air should not be restricted to the
mechanical elimination of large particles, but should also include the removal of biological
pollutants like viruses and bacteria [19–22]. Indeed, infections can be transmitted via the
diffusion of viral particles by air-droplets, through moist soil, surfaces, and water [23]. The
diffusion of viral material in household occupied by infected persons is of particular concern. In
some instances, infection can be propagated by the ingestion of contaminated food, such as
vegetable or fruits, or by drinking non-sanitized water. Viruses may reach the surfaces of fruits
and vegetables when they are fertilized with non-disinfected wastewater. If the contamination
through infected water and food is at the moment well managed in developed countries, the more
difficult access to alimentary resources in developing or underdeveloped areas still represents a
serious public-health issue, seriously undermining human health development.
Furthermore, the quantification of virus content, i.e. the precise determination of the type and
concentration of virus is far from being an innocent subject, and many important scientific
breakthroughs have been realized in the last years. Virus quantification is fundamental not only
7
for the analysis and purification of effluent or wastewater but also for the production of viral
vaccine, antiviral agents, or recombinant proteins. Historically, the more common virus
quantification methods included plaque titer assay, fluorescent focus assay (FFA), and 50 %
tissue culture infection dose (TCID50). Without entering into too much details, that will be out of
the scope of the present review, those methods are based on microbiology techniques in which
hosts cells are exposed to viral samples and the number of infected cells that will lyse eventually
identified by immunostaining, is recorded giving an indication of the virus quantity. The actual
quantification in this case is achieved providing either the number of plaques formed by the
lysed cells, or of the stained foci, expressed per unit of sample and volume, or by its logarithm
(log10). Conversely, TCID50, that yields the amount of virus needed to lyse half of the host cell, is
mostly used in clinical frameworks. More recently, direct measures of the amount of viral
material have emerged testing either specific viral protein or its genetic material. In particular,
quantitative polymerase chain reaction (qPCR) is used to amplify and detect the amount of viral
DNA or RNA, while enzyme-linked immunosorbent assays (ELISA) are sensible to the viral
proteins and acts through an antigenic mechanism. It is worthwhile to mention that the use of
transmission electron microscopy (TEM) may allow to quantify some virus via direct
visualization. In this case, the quantity of virus is expressed directly as the quantity of viral
material, or particles, per unit of sample and volume. Modern methods such as qPCR or
antigenic tests are much faster than the classical cellular based techniques, and have the
advantage to directly target the viral material, although they may be subjected in some instance
to contamination [17,24].
The presence of viruses in air and water can be faced using different technologies with a varied
degree of sophistication, operational costs, removal efficiency and energy consumption. These
technologies comprise mainly mechanical and electrical filtration using for example ceramic
filters [25–27], coagulations processes [28], saturated soil column [29], reclamation systems
8
[30–32], membrane bioreactor process [33], photocatalytic disinfection [34,35], concentration
methods [36], ozone generators and UV irradiation [37], chemical disinfection and membrane-
based processes [38]. Among these technologies, membrane-based processes based on
adsorption appear as a low cost and efficient technique.
Different membranes have been proposed for virus removal from water and air: ultrafiltration
[39,40], nanofiltration [41] and reverse osmosis membranes [20,42–45]. Nowadays, a large
variety of materials has been proposed to this aim, also exploiting specific interactions with the
viral capsid. They include activated carbon [46–49], polysaccharide (PS)-based materials [50],
kaolinite and fiberglass [51], Cu and Ag compounds [52,53], quartz sand [54], functionalized
chitosan nanofibers [55], hydrochar [56], nano-TiO2 [57], resins [13], poly(ethylenimine) [58],
Fe/Ni nanoparticles [59] and zeolites [60]. Alternative strategies for fluid decontamination can
rely on the treatment of wastewaters with natural occurring or biocompatible organic
compounds, such as carbohydrates, that show high antiviral effect coupled with the absence of
unwanted side effects such as cytotoxicity. Note that the majority of these studies was focused
on the removal of the viruses from water but the results can be extended and used as a basis to
develop purification methods for air.
The present review will focus on the different strategies for the rational design of adsorbents
materials that can be easily implemented to eliminate viral particles from air or liquid fluids.
After describing in details the structure of viruses, this review will be organized in sections
corresponding to the main classes of adsorbents: carbon-based materials, oxide-based materials,
zeolites, silica, metal–organic frameworks, clays and carbohydrates. We will present and
critically comment the available practical technologies for removing viruses by adsorption and
our perspectives will be provided in the last section of this review.
9
2. Description of virus structures
In addition, to the possible threats to public health system and social organizations, virus biology
is also complex and fascinating. By a general definition, viruses are small microorganisms that
do not possess a cellular structure. Overall, the classification of viruses may be diverse and
rooted either on molecular biology or phenomenological and phenotypical factors such as: 1) the
type of nucleic acid encoding the genetic information, 2) its structure (single- or double-chain,
linear, circular, fragmented, unregimented), 3) the structure, dimensions, type of symmetry and
number of capsomers (i.e. the subunits constituting the virus external shell, termed capsid, and
mainly constituted of proteins), 4) the presence or absence of the outer shell (super capsid), 5)
the antigenic structure, 6) the genetic interactions taking place, 7) the geographical distribution,
8) the route of infection transmission and 9) the type of host (animal, plant, bacteria) [4,61,62].
Based on the literature, the best known kind of viruses are: rotaviruses, adenoviruses, hepatitis
type A, B, C, poliomyelitis (poliovirus) [25] and coronaviruses. The size of the virus structural
units (virions) range from 20 to 300 nm [20]. The composition of virions includes nucleic acids,
either ribonucleic (RNA) or deoxyribonucleic (DNA) acids, enclosed by a membrane and a
protein shell constituting the capsid, see Figure 1. Protein capsid protects and encloses the viral
genome encoded by the nucleic acids. Thus, protein capsid and nucleic acids are globally known
as nucleocapsid. Note that at a microscopic level the capsid, and hence viruses, may assume
different shapes: cubic, spherical, rod-shaped, etc. [62]. The specific shape of the capsid and the
structural proteins defining the interactions with the cellular receptors will determine on the one
side the type of organisms and cells that can be infected by a given virus, and on the other side
the transmissibility and the possibilities for the interactions with external materials. For instance,
small size viruses can be diffused further with fluid emission and through air compared to
heavier and bigger structures.
10
Figure 1. Illustration of a virus structure.
The most widespread classification of viruses depends on the nature and structure of their
genome rather than on the diseases that they can cause [3]. As a consequence, a basic difference
exists between DNA and RNA viruses, which may have single or double chains of genetic
material, see Figure 2. Furthermore, single chain RNA viruses are divided into RNAs with
positive polarity and RNAs with negative polarity [3]. Typically, DNA viruses replicate in the
nucleus of the host cell, while the replication process of RNA viruses takes place in the
cytoplasm. At the same time, some positive-polarity single-chain RNA viruses, called
retroviruses, although replicating in the cytoplasm, use a completely different strategy involving
the preliminary retro-transcription of RNA into DNA performed by the viral retrotranscriptase.
A further difference in viral strains can be related to the structural properties of the virions
leading to the definition of simple and complex particles [3].
11
Figure 2. Schematic illustration of A) influenza A virus (i.e., RNA virus) and B) hepatitis B
virus (i.e., DNA virus) [4].
Simple viruses are constituted by an external shell, named capsid, formed by the specific
external proteins named capsomers, and whose shape determines the virus symmetry. Complex
viruses have an additional outer shell, the supercapsid, located on top of the capsid. The
supercapsid contains the inner protein layer, formed by the M-protein, and a more bulky layer of
lipids and carbohydrates, that is formed by molecular constituents extracted from the host cell
membranes [4]. In addition, viral glycoproteins penetrate into the supercapsid resulting in curly
protrusions (spikes, fibers) and act as receptor allowing the interaction with cell membrane
proteins and hence the infection [3]. Examples of different classes of viruses are illustrated on
Figure 3.
12
Figure 3. Examples of viruses: A) Adenovirus, B) Sulfolobus turreted icosahedral virus I and C)
Mammalian orthoreo virus [62].
Although viral particles may, on a first approximation, be regarded as structured objects,
traditional separation methods like adsorption and capture have shown a generally low
efficiency, when used in practical applications. Probably, the reason for this lack of performance
may be associated to the specific physicochemical and biochemical properties of viral particles,
namely, their low diffusion mobility, that can be traced back to their molecular size, and the
complex multicomponent and rather labile structure of the virions [63,64]. Indeed, the structural
properties of the virus’ surface determine the mechanism of their interactions with the adsorbent
surface, its bio-affinity, and ultimately the adsorption rate, which is usually accompanied by the
formation of highly specific (or bio-affinity) ionic, hydrophobic and coordination bonds
[25,65,66]. Thus, the physicochemical properties of viruses such as their size, morphology,
surface charge, capsid conformation and the distribution of charged, hydrophilic, and
hydrophobic amino acids strongly affect the virus-material interaction [67,68] as schematically
reported in Figure 4.
13
Figure 4. Capsid structure and hydrophilic/hydrophobic amino acid distribution of MS2 and
PhiX174 phages (bacteriophage). Adapted from [67] with permission from Elsevier.
Hence, the study, also at molecular level, of the specific interplay between the adsorption
processes and the specific nature of the viral envelope is essential for the development and the
improvement of separation methods and strategies for viruses capture and elimination. In
addition to structural optimization, mainly based on the adaptation of the adsorbent material to
the virus size [3], surface charged material could be used to exploit electrostatic effects to
enhance viruses adsorption and inactivation [26,69,70]. Some specific viral particles have been
extensively used to model the adsorption of viruses since they permit a sufficiently large
representation of the different structural and physico-chemical space spanned by the different
pathogens and, as reported in Figure 5, they include the bacteriophages male specific type 2
(MS2), fr, GA and Qβ.
14
Figure 5. Structures of most common virus-like particles – bacteriophages MS2, fr, GA and
utilized as models in adsorption experiments. Reprinted with permission from [66].
15
3. Carbon-based materials
Carbon-based materials are the most used adsorbents for water and air treatment and,
consequently, they have been also largely applied in virus removal. A summary of different
carbon-based materials used for virus adsorption is reported in Table 1 and critically analyzed in
the following. Two types of activated carbon, conventional granular activated carbon and an
activated carbon fiber composite, have been tested and applied for virus removal from water
[48]. The raw material was characterized by a rigid mass of interlocked fibers of an average
length of 0.1-0.4 mm and an average width comprised between 5 and 100 µm. The bacteriophage
MS2 (having a radius of approximatively 25 nm) was chosen as a model for this study. It has to
be noted that bacteriophage is easier and cheaper to grow and test, especially compared to viral
particles and, consequently, it is often used as a model virus especially to study water sanitation.
It was shown that the shape of activated carbon could either inhibit or enhance the removal of the
large bacteriophage particles. Indeed, the study confirmed that carbon fiber composite having a
lower total area (840 m2/g) was more efficient for virus adsorption than granular activated
carbon with larger total area (1050 m2/g) resulting in a higher virus removal due to different
shape and size fraction of the activated carbons.
Matsushita et al. [47] investigated the removal of the two bacteriophages (diameter 23.5 ±
0.8 nm) and MS2 (diameter 22.5 ± 1.0 nm) by adsorption on commercially available powdered
activated carbon (N-PAC) and super-powdered activated carbon (S-PAC). This study allowed to
pinpoint some crucial factors for virus removal, namely: (i) increasing the electrophoretic
repulsive force between virus and particles of powdered activated carbon results in the
decreasing of virus removal, (ii) the increase of N-PAC pore volume improved the virus
removal, a pore diameter of 20-50 nm was shown to be necessary to trigger adsorption, (iii) the
increase in hydrophobicity of the virus surface results in a higher removal, and (iv) the low
negative surface charge of AC surface enhanced the virus removal. Similar findings were
obtained by Nemeth et al. [72] who proposed multi-walled carbon nanotubes (MWCNT)
16
sensitized by Cu2O, TiO2 and Fe2O3 NPs in order to build hybrid membranes for MS2
bacteriophages removal from contaminated water. Cu2O-coated MWCNT membrane (see Figure
6a) showed the highest activity in virus removal in the pH range comprised between 5.0 to 9.0,
yielding a 4-Log, i.e. 99.99 % lower bound, for the percentage of virus retention, and,
consequently, were considered promising for virion uptake (Figures 6b and 6c).
Table 1. Summary of some advantages and limitations of carbon-based materials used for virus
inactivation.
Adsorbents
Type of virus
Consequences and advantages
-Granular activated carbon (GAC);
-Activated carbon fiber composite
(ACFC)
-Bacteriophage MS2
-ACFC adsorbent was more efficient
for virus adsorption than GAC;
-Due to its shape, GAC was not the
best adsorbent for virus besides both
surface areas are comparable
-Commercially powdered activated
carbon (N-PAC);
-Super-powdered activated carbon (S-
PAC)
-Bacteriophages
-MS2
-Three factors enhanced the
adsorption of virus by these
adsorbents:
1-A large pore diameter of 20-50 nm;
2-The increase in hydrophobicity of
the virus surface;
3-A low negative surface charge of
activated carbon;
-Increasing the electrophoretic
repulsive force between virus and
particles of powdered activated carbon
results in the decreasing of virus
removal
-MWCNT sensitized by Cu2O, TiO2
and Fe2O3 NPs
-MS2 bacteriophages
-Cu2O-coated MWCNT membrane
showed the highest activity in virus
removal at pH 5 -9;
-CuO does not improve the virus
removal
-TiO2 and Fe2O3 NPs- MWCNT have
not significantly enhanced the virus
removal
17
Figure 6. (a) TEM micrograph of Cu2O/MWCNT nanocomposite membrane with 25 % of
MWCNT, (b) MS2 bacteriophage removal using MWCNT-based nanocomposite membranes in
batch experiments and (c) MS2 bacteriophage removal using MWCNT-based nanocomposite
membranes in flow experiments. Adapted from [72].
4. Oxide-based materials
It is known that viral capsid surfaces are negatively charged, hence surface positively charged
nanoparticles may be well suited for virus removal through electrostatic interaction, at least at
suitable pH [73]. Indeed, electrostatic-mediated adsorption inactivates viruses and can be used
both for water and air disinfection. Table 2 provides a summary of oxide-based materials
employed to remove or inactivate viruses. Mazurkow et al. [71] reported spray-dried alumina
18
granules modified with copper (Cu) and copper oxide (Cu2O) nanoparticles (NP) as effective
adsorbents against viruses, thanks to the capacity of the nanoparticles to create a high density of
adsorption sites. The experiments were conducted in flow reactor and showed a removal of 99.9
% of the MS2 bacteriophages by Cu2O and Cu, conversely CuO had no impact on virus removal.
Gutierrez et al. [73] performed batch and dynamic experiments in order to investigate the
removal of group A porcine rotavirus (whose diameter is of 74.57 ± 1.32 nm) and bacteriophage
MS2 (diameter of 25.42 ± 0.93 nm) using glass fiber coated with hematite Fe2O3 NPs having a
surface area of 80.75 m2/g and total pore volume of 0.0835 cm3/g. The batch mode disinfection
demonstrated a high removal of both rotavirus and bacteriophage MS2 quantified by 2.49 × 1011
plaque forming unit/g and 8.9 × 106 focal forming unit/g, respectively (Figures 7a and 7b), while
the concentration of hematite used was 0.043 and 0.26 g/L for virus and bacteriophage MS2,
respectively. Transmission electron microscopy (TEM) images confirmed the electrostatic
adsorption of MS2, while they also underlined structural damages to the rotavirus caused by the
interaction with hematite that also participate in the virus inactivation. (Figures 7c and 7d).
While hematite NPs present the advantage of offering more available adsorption active sites
coupled with strong electrostatic attraction, which can be beneficial in favoring virus depletion, it
was concluded that the inactivation of viral particle might depend on the specific proteins of its
capsids or the robustness of its structure [73], and hence its efficacy could vary considerably
between different viral strains. However, it is clear that hematite NPs may represent a most
valuable system to achieve virus removal from contaminated water and its use could be extended
to air treatment.
Table 2. Summary of some advantages and limitations of oxide-based materials used for virus
inactivation.
19
Adsorbents
Type of virus
Consequences and advantages
-Spray-dried alumina
granules modified with
copper (Cu) and copper
oxide (Cu2O)
nanoparticles (NP)
-MS2 bacteriophages
-A removal of 99.9 % of the MS2 bacteriophages by Cu2O and Cu;
-Glass fiber coated with
hematite Fe2O3NPs;
-Biosand filter filled with
iron oxide NPs
-Bacteriophage MS2;
-Rotavirus
-A high removal of both viruses;
-NPs have the advantage to provide more available adsorption sites
coupled with strong electrostatic attraction leading to virus depletion;
-Hematite NPs may represent a most valuable system to achieve virus
removal from contaminated water and its use could be extended to air
treatment;
-A significant higher removal of bacteriophage MS2 by an iron-
sensitized sand column compared to a sand-only column;
-The removal mechanism was explained by the electrostatic interaction
of negatively charged MS2 particles with the positively charged iron
oxides;
-It was concluded that the inactivation of viral particle might depend on
the specific proteins of its capsids or the robustness of its structure, and
so its efficacy could vary considerably between different viral strains
Magnesium oxyhydroxide
-Bacteriophages MS2
-PhiX174
-A favorable electrostatic interactions between the negatively charged
virion particles and the positively charged patches of magnesium
oxyhydroxides was improved the virus adsorption;
-Magnesium oxyhydroxide could be a good adsorbent for viruses also
due to its availability, high isoelectric point and low cost
-Diatomaceous-based
ceramic filter
-Enterobacteriaphage
MS2;
-Enterobacteria phage
PhiX174;
-Wild-type
bacteriophage
Siphovirida
-The experimental conditions were not optimized for virus removal;
-It is not very suitable for virus removal due to existence of repulsive
forces between the viruses and the filter surface besides steric clashes
caused by phage specific knobs.
Nano-TiO2-membranes;
Spherical Fe/NiNPs
-Bacteriophage f2
-Bacteriophages f2
-It is an effective adsorbent to remove the virus;
-Fast removal (after 60 min);
The virus removal was governed by the electrostatic attraction between
the opposite charged nano-TiO2 and phage f2 under acid conditions;
-Bacteriophages f2 were completely removed by Fe/Ni NPs after 30
minutes
Silica-decorated TiO2
NPs
-Bacteriophage MS2
-It was shown that doped-TiO2 including 5 % of SiO2 has a 37-fold
enhanced viral removal performance compared to pristine TiO2;
-It was suggested that this removal approach has the advantage of using
low-cost and green materials.
20
Figure 7. Removal of MS2 and rotavirus onto hematite NPs surface: (a) MS2 removal by
hematite NPs; (b) rotavirus removal by hematite NPs; (c) TEM image of MS2 bacteriophage
adsorbed onto hematite NPs surface; (d) TEM image of structurally damaged rotavirus particles
contacted with hematite NPs surface. Reprinted from [73].
Similarly, Bradley et al. [74] have studied the potential application of a biosand filter filled with
iron oxide NPs to remove virus, see Figures 8a and 8b. The iron oxides were obtained as result
of zero-valent iron particles corrosion, while the purification experiments were performed in
21
continuous flow process and showed a significantly higher removal of bacteriophage MS2 by an
iron-sensitized sand column (5log10) compared to a sand-only column (0.5log10) (Figure 8c). The
removal mechanism was again explained by the electrostatic interaction of negatively charged
MS2 particles with the positively charged iron oxides. In this same line of investigation,
Domagala et al. [75] used Cu2O/MWCNTs filters to remove MS2, finding that while the filters
were efficient, electrostatic interactions should not be held as the sole responsible of the
adsorption [75].
Figure 8. (a) Breakthrough curves of NaCl tracer from glass columns packed with sand and
zero-valent iron; (b) small-scale glass columns packed with iron oxides used for virus removal;
(c) the log10 reductions bacteriophage MS2 obtained by continuous flow through clean quartz
sand with no iron particles and sand mixed with 10 % (vol.) iron. Reprinted from [74].
Michen et al. [26] have proposed the use of magnesium oxyhydroxide, embedded into a ceramic
filter (10, 15, and 20 % of MgO), to improve viruses removal from water. These filters showed
an enhanced removal of about 4-log of bacteriophages MS2 and PhiX174. Once again, the
enhanced removal of the virus was attributed to the establishment of favorable electrostatic
interactions between the negatively charged virion particles and the positively charged patches of
22
magnesium oxyhydroxides. Furthermore, it was also concluded that viruses may be inactivated
in an alkaline environment and the presence of the slightly soluble Mg(OH)2 can indeed cause
the increase of water pH above 9. Albeit efficient for viral decontamination, the high increase of
pH alters the physico-chemical properties of water and, in particular, makes it not suitable for
drinking purposes; hence, a continuous flow mode was proposed for use of the ceramic filters.
Globally, the results suggested that magnesium oxyhydroxide could be a good adsorbent for
viruses also due to its availability, high isoelectric point and low cost.
The finding that viruses can be attracted by positively charged surface has been confirmed by
Wegmann et al. [42], in analyzing the impact of electropositive zirconium (hydr)oxide NPs
coatings on the performance of ceramic filters to remove viruses from water. Indeed, the internal
filter surface was coated by ZrO2 nanopowder via dip-coating and thermal treatment, resulting in
a considerable increase of the specific surface area of the filters going from 2 to 25.5 m2/g after
coating. The virus removal rate was also significantly increased, most notably achieving a 7-log
removal, equivalent to 99.99 % of the MS2 bacteriophages. Thus, zirconium oxide coating was
suggested as promising for virus filtration techniques with potential applications in both water
and air treatments.
On the other hand, the removal of positively charged viruses through negatively charged
surfaces has been investigated by Michen et al. [25]. Three bacteriophages, Enterobacteria phage
MS2 (isoelectric point (IEP) of 3.5, diameter 25 nm), Enterobacteria phage PhiX174 (IEP 6.6,
diameter 26 nm) and wild-type bacteriophage Siphovirida (IEP 2.7, diameter 60 nm), were used
as model viruses. The results have been rationalized through the use of the DLVO
(Derjaguin−Landau−Verwey−Overbeek) theory [25]. Most notably, it was shown that
diatomaceous-based ceramic filter, which is widely used to purify water, was not suitable for
virus removal due to repulsive forces between viruses and the filter surface besides steric clashes
caused by phage specific knobs.
23
Zheng et al. [57] reported the application of nano-TiO2-membranes for phage f2 removal from
water, see Figures 9a and 9b. This study showed that the removal of phage f2 was effective after
60 min and the results have been rationalized with the Freundlich adsorption model yielding a qe
= 27.4.Ce1.24. PolyVinyliDene Fluoride (PVDF) (0.20 μm) and PolyAcryloNitrile (PAN) (0.05
μm) membranes were used for TiO2 coupling in order to use nano-TiO2 membranes in flow
reactors for water treatment (Figures 9c and 9d). The experiments demonstrated higher virus
removal by PAN (3.88-log) compared to PVDF membrane (6.40-log). The mechanism of virus
removal was explained by the electrostatic attraction between the opposite charged nano-TiO2
and phage f2 under acid conditions. The coupled TiO2/membrane system was stable and can be
used for virus removal from contaminated water. Silica-decorated TiO2 NPs have been used by
Liga et al. [76] for the bacteriophage MS2 inactivation in water. It was shown that doped-TiO2
including 5 % of SiO2 has a 37-fold enhanced viral removal performance compared to pristine
TiO2. The authors suggested that this removal approach has the advantage of using low-cost and
green materials and, hence, is particularly appealing for water treatment.
24
Figure 9. (a) nano-TiO2 membrane adsorption reactor; (b) adsorption kinetics of nano-
TiO2 regarding phage F2. SEM images of (c) PVDF (PolyVinyliDene Fluoride) and (d) PAN
(PolyAcryloNitrile) flat membranes. Reprinted from [57].
Cheng et al. [59] reported the removal of bacteriophage f2 from water by spherical Fe/Ni
NPs (93 nm) under aerobic conditions, see Figure 10. Fe/Ni NPs were synthesized by a liquid-
phase method using NaBH4 solution. It was found that bacteriophages f2 were completely
removed by Fe/Ni NPs after 30 minutes treatment from a solution with an initial concentration of
4×106 PFU/mL (PFU: plaque-forming unit). The effect of the Fe:Ni ratio on virus removal was
investigated and results showed that the highest removal efficiency was obtained with a 5:1 ratio.
Oxygen condition, pH, initial virus concentration, adsorbent mass and temperature were also
25
tested and analyzed to determine their impacts on the removal of this virus. The inactivation
mechanism of bacteriophage f2 was explained by the oxidative stress produced by iron and by
the production of reactive oxygen species (OH and O2•−) formed due to the oxidation of Fe0 and
catalyzed by Ni0 [59].
Figure 10. Removal of bacteriophage f2 in water by Fe/Ni nanoparticles. Reprinted from [59].
5. Zeolites
Zeolites are crystalline and porous materials with regular framework structures composed of
cages and pores of various size and shape [77]. Zeolites have exceptional chemical selectivity,
high adsorption capacity and are biocompatible and safe, consequently, they have attracted
considerable interest for biomedical applications. It is now possible to synthesize zeolites with
nanosized dimensions and stabilized in colloidal suspensions [78,79]. The high stability of
nanozeolites with regular micropores makes them suitable for selectively adsorbing and
desorbing different molecules based on their size allowing them to act as molecular sieves [80].
Furthermore, the tunable chemical composition of these crystalline materials allows them to
adapt to different chemical environments and thus to exhibit high stability in acidic and alkaline
media [81]. The interest of zeolites as antiviral adsorption agents also resides in the fact that their
properties can easily be tuned by the introduction of different metal cations in nanosized
26
materials, most notably altering their adsorption capacity [80,82–85]. Although this has mainly
be exploited for gas separation, its interest for purification stands out as well. Zeolites can also
be specifically functionalized, tuning their surface and interparticle porosity, in order to target a
specific tissue or cell type. As an example, antibody-nanozeolite bioconjugates have been
generated and successfully employed to target cancer cells [86]. As an extension, and because of
their favorable physico-chemical properties zeolites can be efficiently used to remove viruses,
also thanks to their high porosity and surface area (see Table 3).
Bright et al. [60] used zeolite as an adsorbent to remove coronavirus 229E from water. They
demonstrated that silver- and copper-doped zeolites can induce a significant reduction of the
presence of the coronavirus in water solution after 1 h treatment. Indeed, the functionalization of
zeolite via silver and copper, i.e. metals presenting known antibacterial properties, has been
shown to be an efficient way to inhibit SARS coronavirus, as well as other coronaviruses and
human norovirus (calicivirus) [60]. Even if experiences are still scarcer than for other materials,
the promising results show that metal-functionalized zeolites can be considered as an excellent,
and highly flexible, agent to reduce the contamination by all types of viruses from both water
and air.
6. Silica nanoparticles
Silica is a well-known adsorbent that has unique physicochemical properties: large surface area,
chemical purity, hydrophilicity, and significant adsorption capacity. Silica has the ability to
actively adsorb bioactive molecules, proteins, microorganisms, and viruses [36,87–89]. The
specific surface area of silica used for medical purposes is 300 m2/g, and is usually characterized
as a white powder consisting of spherical particles with a diameter of 9-10 nm. The surface of
silica particles contains silanol and hydroxyl groups, which are interacting via hydrogen bonds
[90]. In addition, water molecules are present at the silica surface both chemically bonded or
27
physically adsorbed. The concentration of silanol groups, which are the main adsorption sites, is
approximately 0.6-0.7 mmol/g or 2-2.5 μmol/m2 [91,92].
28
Table 3. Summary of some advantages and limitations of different alternative adsorbents used
for virus inactivation.
Adsorbents
Type of virus
-Zeolite;
-Novel amine-
functionalized
magnetic
Fe3O4–SiO2
NH2 NPs
-Coronavirus 229E
-SARS coronavirus;
-Bacteriophage f2
and Poliovirus-1
-Silica
nanoparticles:
A variety of
NH2-containing
moieties can be
used to modify
silica, for
example,
lupamin, 3-3-
(ethylenediami
no) propyl,
aminopropyl,
3-
(diethylenetria
mino) propyl;
-Novel amine-
functionalized
magnetic
Fe3O4–SiO2
NH2 NPs
-Bacteriophage MS-
2;
-Bacteriophage f2
and Poliovirus-1
-Metal–Organic
Frameworks
(MOFs)
-Zika, Dengue,
human immune
deficiency virus-1
(HIV-1) and
Japanese
encephalitis virus
29
-Clay and clay
minerals
-Rotaviruses and
coronaviruses
-Cyclodextrins
( CDs)
-Herpes simplex
virus (HSV);
-Respiratory
syncytial virus
(RSV);
-Dengue virus, and
Zika virus
Silica has a significant protein adsorption capacity, for example, 1 g of silica can adsorb up to
200-300 mg/g of gelatin or 800 mg/g of albumin from aqueous solution at pH 5-6. Silica is an
active adsorbent of various kind of microorganisms: indeed 108 up to 1010 microbial bodies can
be adsorbed per 1 g of silica. The properties and structure of the active centers of the SiO2
surface, which determine the specific interaction with the biological objects, are described in
[93–96].
As such silica-based NPs are particularly attractive for virus removal as highlighted in Table 3.
Various methods of SiO2 surface modification have been described in order to increase its
adsorption selectivity towards microorganisms [97–101]. Chemically modified silicas are of
particular interest among nanoporous mineral oxides [102]. They have a developed surface and
are not only excellent adsorbents in their original form, but they are also capable of radically
changing their physicochemical properties as a result of chemical modification [103]. A variety
of NH2-containing moieties can be used to modify silica, for example, lupamin, 3-3-
(ethylenediamino) propyl, aminopropyl, 3-(diethylenetriamino) propyl [104]. Such modified
functionalized silica particles show high efficiency towards bacteriophage MS-2 removal (> 98
%). The mechanism of virus inactivation is explained by the high density of positively charged
NH2-groups, electrostatic interaction and hydrogen bonds formation.
30
As another example, magnetic silicon microspheres have been used for detection of Japanese
encephalitis virus [105]. Novel amine-functionalized magnetic Fe3O4–SiO2–NH2 NPs have been
investigated as adsorbent for bacteriophage f2 and Poliovirus-1 removal by Zhan et al. [106].
NPs were obtained through layer-by-layer method using Fe(acac)3, tetraethyl orthosilicate
(TEOS) and γ-amino propyltriethoxy-silane (APTES) as precursors. The adsorption mechanism
was rationalized by the surface charge, hydrophobicity and surface properties of pathogens and
NPs matrix. The electrostatic attraction between the negatively charged Poliovirus-1 walls and
the positively charged adsorbent surface was clearly identified, and it was shown that the
removal efficiency of Fe3O4–SiO2–NH2 NPs over bacteriophage f2 and Poliovirus-1 exceeded 97
%. Cademartiri et al. [107] have explored an alternative strategy in which four unmodified
phages were immobilized on anionic silica substrate via electrostatically-facilitated
physisorption, see Figure 11. The proposed adsorbents can be used to detect and control
pathogens, particularly drug resistant bacteria.
Figure 11. Electrostatic attraction model between phage and charged SiO2 particles. A: tail down
and non-specific adsorption, B: head down, C: head down saturated surface [107].
7. Metal–Organic Frameworks (MOFs)
31
MOFs are composed of inorganic nodes of metallic species/clusters and organic ligands. These
materials are very stable under different operating conditions (e.g., temperature and solvents)
[108]. Several studies have concluded that they offer a spectrum of possibilities to prepare new
materials including biocomposites [108,109]. The application of MOFs-based biocomposites
covers biocatalysis, DNA detection, sensing, drug delivery and immobilization [109]. It has been
shown that the textural properties (pore size and volume), crystal morphology and chemical
functionality of MOFs can be modulated to obtain specific performances for a variety of
applications [110,111]. MOFs properties can be tailored to encapsulate proteins and other
biological molecules preserving their functionalities [109]. These materials are also capable of
encapsulating bacteria and viruses in which case adsorption plays a fundamental role [108].
These findings open a broad spectrum of possibilities to employ MOFs for facing the
environmental and public health challenges caused by viruses, see Table 3.
The analysis and study of biological-inorganic matrixes obtained from MOFs are relevant topics
to improve and enhance the environmental and medical applications of this new type of
biomaterials. The organic functionalities of MOFs play an important role in the anchorage of
biomacromolecules for biocomposites synthesis. For instance, Doonan et al. [109] discussed and
analyzed the mechanisms of the incorporation of biological compounds on MOFs, which can be
performed via encapsulation, infiltration or bioconjugation (Figure 12). Note that MOFs
structures can host a wide variety of both inorganic and organic compounds including biological
macromolecules. Parameters like the composition, topology, porosity and surface functionalities
of MOFs are important and should be analyzed during the synthesis of biocomposites and to
perform the coating of living systems (i.e., biomimetic mineralization) [108].
However, it has been recognized that the incorporation of biomacromolecules on the MOFs pore
structures implies several technological challenges. Previous studies have indicated that
biomacromolecules encapsulated on MOFs can be released by changes in the medium
32
conditions, for example, pH [109]. On the other hand, MOFs can be used in vaccine
developments based on the fact that these materials can coat and work as exoskeletons to protect
living cells from aggressive environments [112]. Ricco et al. [108] have concluded that MOF can
be used to improve the cold chain with the aim of avoiding the degradation of drugs and vaccine.
It has been also suggested that hybrid biosynthetic materials can be prepared from MOFs to
improve the pharmacokinetics of viral nanoparticles thus affecting the immune response [108].
Vaccine design based on MOFs is a promising area that however still needs to be much largely
explored and studied to reach commercial applications.
Finally, MOFs-based materials have also been used to develop sensor for detecting Zika
[113,114], Dengue [113], human immune deficiency virus-1 (HIV-1) [115] and Japanese
encephalitis virus [116] with promising results.
Figure 12. Mechanisms for the incorporation of biological compounds on MOFs structures.
Reprinted from [108].
8. Clay and clay minerals
33
A variety of promising adsorbents obtained from clays and their composites has also been
proposed to remove rotaviruses and coronaviruses [105], see Table 3. Indeed, the diameter of
rotavirus and coronavirus particles are 60-80 and 60-220 nm [117], respectively, which is clearly
larger than the size of pores of different adsorbent particles. Hence, both viruses can be adsorbed
on the outer surface of adsorbents. The experimental results demonstrated that clay-based
adsorbents showed good (70-90 %) to excellent (> 90%) capability to remove bovine rotavirus
and bovine coronavirus [117]. Based on this investigation, it was suggested that the involved
interactions between the adsorbent surfaces and viruses is probably due to a non-specific protein
binding.
9- Viruses adsorption by carbohydrates and their derivatives
Several biological molecules, carbohydrates in particular, have shown effective antiviral
activities; these materials are also summarized in Table 3. Such activity is usually due to the
competition with the host cell for the interaction with viral capsid proteins, hence preventing the
first step of cell infections. Some of the most efficient agents in this context mimic the sugars
present at the cell surface, in particular heparan-sulfate proteoglycans that are responsible for the
initial viral attachment [118–120]. The chemical structure of these carbohydrates, their origins
and their antiviral properties are briefly described below.
9.1. Antiviral effect of sulfated polysaccharides
Various polysaccharides from plants, marine vegetables, algae and lichens have received
increasing attention due to their ability to scavenge free radicals, to activate immune systems, to
inhibit lipid peroxidation and to inhibit viral replication [121]. Consequently, these
polysaccharides have exhibited numerous biological activities including anticoagulant, antiviral,
antioxidative, anti-inflammatory, antiangiogenic, antithrombotic and anticancer effects [122–
127]. Sulfated polysaccharides (SPs) especially from sea algae such as fucoidans, carrageenans,
34
heparin, neem, laminarans and ulvansare are considered as the most interesting and attractive
carbohydrates showing antiviral properties [128–133].
Since 1978, Richards and coworkers [134] have recognized the inhibition of viruses by
polysaccharide fractions from marine algae. Recently, other studies have reported that several
plants and marine algae contain significant quantities of complex sulfated polysaccharides that
can exhibit antiviral activity against serious infectious diseases such as HIV [135], yellow fever
virus [136], dengue virus type 1 [136], dengue virus type 2 [137] and herpes simplex viruses
[133,138–140].
A number of hypotheses has been proposed to explain the antiviral mechanisms of sulfated
polysaccharides (see Figure 13) including direct interaction with the viral particle, alteration of
viruses’ adsorption or attachment to cell receptors, inhibition of the penetration of virus into the
host cell, the interference with different stages of viral replication [141–143] .
35
Figure 13. Antiviral action mechanisms of medicinal plant metabolites. Reprinted from [143].
On the other hand, numerous studies have reported efficient strategies, based on synthetic
pathways, enzymatic digestion, partial hydrolysis and methylation methods, to produce sulfated
derivatives of different polysaccharides such as dextran, pentosan, dermatan, carrageenan,
alginate, xylomannan and galactan [123,135,144]. Furthermore, it has been highlighted that the
presence of sulfated groups enhanced the antiviral properties, and is hence essential for viruses’
inactivation. More specifically, it has been noted that while the virucidal efficacy can increase
with the degree of sulfation and the specific position of the sulfate ester groups, it also depends
on the molecular weight, the constituent sugars, their conformation and dynamic stereochemistry
[132,145,146]. Generally, the results have indicated that the greater the degree of sulfation, the
higher the biological activity, an observation which could conduct to find a drug candidate with
higher potency and less cytotoxicity [141,147,148], since polysaccharides with low sulfation
36
degrees are in general inactive against viruses [145]. Some applications of these natural
compounds in virus inactivation have already been reported [121,137]. For example, Hidari et al.
[137] have found that dengue virus particles bounded exclusively to fucoidan via their envelope
glycoprotein thus indicating that sulfated polysaccharides could be developed as a potential
inhibitory agent. Kim et al. [121] reported that marine sulfated polysaccharides presented
numerous advantages such as relatively low production costs, broad spectrum of antiviral
properties, low cytotoxicity, safety, wide acceptability and novel modes of action over other
classes of antiviral drugs. Hence, they suggested the use of these sulfated polysaccharides as
promising decontamination agents in the near future.
9.2. Antiviral action of modified and sulfated cyclodextrins
Cyclodextrins (CDs) are natural cyclic oligosaccharides formed by α-(1-4)-linked glucose units
[149,150]. The most common CDs, referred as α-, β- and γ-CDs, have 6, 7 and 8
glucopyranoside moieities, respectively. These cyclic natural compounds have attracted a
considerable attention especially for their biological applications due to their ability to
encapsulate different compounds being, consequently, used as molecular carriers [151]. Thus,
CDs have found numerous applications in commercial and industrial fields including drug
delivery [151,152], gene delivery [153], bioimaging [154], photodynamic therapy [153], air
fresheners [153], cosmetics [153], food processing [155] and environmental depollution
[156,157]. As molecular carriers, CDs have been employed to enhance the antiviral activity of a
phosphodiester oligodeoxynucleotide [158]. This study showed that an important increase of the
antiviral activity (90 % inhibition) was obtained with only 7.5 mM oligonucleotide complexed to
a cyclodextrin derivative, the 6-deoxy-6-S-b-D-galactopyranosyl-6-thio-cyclomaltoheptaose, in
the 1:100 ratio. The authors suggested that the use of cyclodextrin derivatives as carrier for
phosphodiester oligonucleotides delivery could be an effective method for increasing the
37
therapeutic potential of these compounds against viral infections. On the other hand, De Clercq
[159] reported that the modified cyclodextrin sulfates acted, by themselves, as anti-HIV agents
since they interfered with the adsorption stage of HIV replicative cycle. It was also concluded
that the presence of sulfate groups was necessary to induce the anti-HIV activity as it is essential
for the inhibition of the virus-cell binding. CDs chemical modifications, like those reported for
the modified β-cyclodextrin sulfates (mCDS71 and mCDS11), is a promising approach to
achieve high oral bioavailability and absorption [160–162]. Further studies have also confirmed
the antiviral properties of sulfonated CDs against HIV [163,164]. However, their action was
found to be virus specific and reversible. Recently, Jones et al. [120] have prepared cyclodextrins
with mercaptoundecane sulfonic acids, to mimic heparan sulfates and obtained outstanding anti-
viral properties. They showed that the modified CDs, while being nontoxic and biocompatible,
possessed in vitro broad-spectrum virucidal properties at micromolar concentrations against
several viruses including herpes simplex virus (HSV), respiratory syncytial virus (RSV), dengue
virus, and Zika virus. Their results also indicate that modified CDs were effective ex vivo against
both laboratory and clinical strains of RSV and HSV-2 in respiratory and vaginal tissue culture
models, respectively. Additionally, they were also effective when administrated in mice before
intra vaginal HSV-2 inoculation. Finally, these modified CDs were successful in overcoming a
mutation resistance test that the available anti-HSV drug (acyclovir) failed.
10. Conclusion and Perspectives
This review covers the potential application of a number of adsorbents or inhibitors for the
treatment of water and air polluted by viruses and to the development of potential antivirals
treatment. Overall, the viruses’ properties and classification have been discussed, including the
available adsorbents that can be used for their removal or inactivation. Indeed, while a variety of
adsorbents can be used to remove different viruses present in the environment, their performance
38
is affected by both the structural parameters and their surface chemistry. It is clear from the
literature review that the surface charge of adsorbents is a paramount parameter for the effective
viral inactivation in both air and water, that can be tailored via chemical functionalization
[165,166]. Indeed, the electrostatic potential of adsorbent surface depends on the concentration
and type of functional groups that are exposed at the adsorbent/fluid interface [166]. The
adsorbent surface composition can be modified via the incorporation of a variety of doping
elements thus improving the surface interactions with viruses. Note that the identification of the
best chemical functionalization protocol to enhance the surface properties of a specific adsorbent
and, consequently, its corresponding antiviral activity is paramount to reduce the purification
costs. Also, the analysis of surface potential and antiviral properties of different adsorbents is a
relevant topic to be explored in forthcoming studies.
Globally, it appears that nanoparticles have a tremendous potential to develop effective and low-
cost sanitation systems for air and water purification. On the other hand, biological compounds
like polysaccharides and cyclodextrins may represent interesting alternative to develop novel
adsorbents for the removal of viruses from fluids. The production or extraction of these chemical
compounds, as well as their inclusion in everyday life objects (e.g., plastic surfaces, textiles,
hand gel) represents a most important challenge that the scientific community has to undertake to
fight against important public health threats.
As stated in this review, practical methods applied for virus removal include membrane
filtration, chemical disinfection with chlorine or ozone and ultraviolet disinfection [27,35].
However, these methods may face different economical and/or technical limitations. For
instance, the application of chemicals for water disinfection (e.g., chlorination and ozonation)
has been reduced in some countries due to the generation of toxic byproducts (e.g., haloacetic
acids, trihalomethane) from the degradation or transformation of the disinfectants [27,167].
Membrane technologies are effective for virus removal [27] but their main drawback is
39
associated with their operating costs, which can be very high and hence prohibitive for large
scale use in underdeveloped countries [11]. Previous studies have also reported that some viruses
(e.g., Norovirus) can resist the UV radiation thus limiting the efficacy of this process [168].
Adsorbents already available in the market for water disinfection include traditional filters with
activated carbons doped with silver [169], chemically modified polymers as ion exchange resins
[170] and even MOFs [171]. Overall, available commercial adsorbents may lack an effective
antiviral activity especially for new viruses besides relatively high production costs that are
mainly due to the specific compound used to confer antiviral properties of the adsorbent surface.
Indeed, in some cases they comprise high-added values compound which can be used for other
technological applications thus increasing their demand and, consequently, their market price.
Specific emphasis should then be placed on the development of efficient and low-cost
purification processes that will allow the large-scale treatment of fluids in developing or
underdeveloped countries to avoid the uncontrolled diffusion, or mutation, of dangerous viral
agents resulting in global threats. The production of self-sanitizing objects will be particularly
attractive and will require a considerable effort from the material science community in the
production of efficient, flexible, and miniaturized purifying agents. It is clear, as also shown by
the COVID-19 pandemic, that developing novel adsorbents with outstanding antiviral properties
will be a permanent and continuous necessity to face the presence of known and unknown
viruses in the environment.
CRediT author statement
Lotfi Sellaoui, Michael Badawi,Tetiana Tatarchuk: Conceptualization, Methodology
40
Lotfi Sellaoui, Michael Badawi, Tetiana Tatarchuk, Antonio Monari, , Sonia Jemli,
Guilherme Luiz Dotto,Adrian Bonilla-Petriciolet: Data curation, Writing- Original draft
preparation
Lotfi Sellaoui, Michael Badawi, Tetiana Tatarchuk, Antonio Monari, Sonia Jemli, Adrian
Bonilla-Petriciolet, Guilherme Luiz Dotto, Zhuqi Chen:Visualization, Investigation
Zhuqi Chen: Supervision
Lotfi Sellaoui, Michael Badawi, Antonio Monari, Adrian Bonilla-Petriciolet, Zhuqi Chen
:Writing- Reviewing and Editing,
Acknowledgements
This work was financially supported by the National Key R&D Program of China (No.
2018YFC1901403), the National Science Foundation of China (No. 21671072), the Fundamental
Research Funds for the Central Universities (No. 2019kfyRCPY058), and Chutian Scholar
Foundation from Hubei province.
References
[1] N. Zhang, L. Wang, X. Deng, R. Liang, M. Su, C. He, L. Hu, Y. Su, J. Ren, F. Yu, L. Du,
S. Jiang, Recent advances in the detection of respiratory virus infection in humans, J.
Med. Virol. 92 (2020) 408–417. https://doi.org/10.1002/jmv.25674.
[2] J. Guarner, Three Emerging Coronaviruses in Two Decades, Am. J. Clin. Pathol. (2020)
420–421. https://doi.org/10.1093/ajcp/aqaa029.
[3] W.-S. Ryu, Molecular Virology of Human Pathogenic Viruses, Elsevier Inc., 2017.
[4] I.M. Artika, A. Wiyatno, C.N. Ma’roef, Pathogenic viruses: Molecular detection and
characterization, Infect. Genet. Evol. 81 (2020).
https://doi.org/10.1016/j.meegid.2020.104215.
[5] A.A. Elkholy, R. Grant, A. Assiri, M. Elhakim, M.R. Malik, M.D. Van Kerkhove, MERS-
CoV infection among healthcare workers and risk factors for death: Retrospective analysis
of all laboratory-confirmed cases reported to WHO from 2012 to 2 June 2018, J. Infect.
41
Public Health. 13 (2019) 418–422. https://doi.org/10.1016/j.jiph.2019.04.011.
[6] N. Zhu, D. Zhang, W. Wang, X. Li, B. Yang, J. Song, X. Zhao, B. Huang, W. Shi, R. Lu,
P. Niu, F. Zhan, X. Ma, D. Wang, W. Xu, G. Wu, G.F. Gao, W. Tan, A novel coronavirus
from patients with pneumonia in China, 2019, N. Engl. J. Med. 382 (2020) 727–733.
https://doi.org/10.1056/NEJMoa2001017.
[7] C. Sohrabi, Z. Alsafi, N. O’Neill, M. Khan, A. Kerwan, A. Al-Jabir, C. Iosifidis, R. Agha,
World Health Organization declares global emergency: A review of the 2019 novel
coronavirus (COVID-19), Int. J. Surg. 76 (2020) 71–76.
https://doi.org/10.1016/j.ijsu.2020.02.034.
[8] J.L. Domingo, M. Marquès, J. Rovira, Influence of airborne transmission of SARS-CoV-2
on COVID-19 pandemic. A review, Environ. Res. 188 (2020) 109861.
https://doi.org/https://doi.org/10.1016/j.envres.2020.109861.
[9] S.V. Mohan, M. Hemalatha, H. Kopperi, I. Ranjith, A.K. Kumar, SARS-CoV-2 in
environmental perspective: Occurrence, persistence, surveillance, inactivation and
challenges, Chem. Eng. J. 405 (2021) 126893. https://doi.org/10.1016/j.cej.2020.126893.
[10] G. Medema, L. Heijnen, G. Elsinga, R. Italiaander, A. Brouwer, Presence of SARS-
Coronavirus-2 RNA in Sewage and Correlation with Reported COVID-19 Prevalence in
the Early Stage of the Epidemic in the Netherlands, Environ. Sci. Technol. Lett. 7 (2020)
511–516. https://doi.org/10.1021/acs.estlett.0c00357.
[11] B. Adelodun, F.O. Ajibade, J.O. Ighalo, G. Odey, R.G. Ibrahim, K.Y. Kareem, H.O.
Bakare, A.O. Tiamiyu, T.F. Ajibade, T.S. Abdulkadir, K.A. Adeniran, K.S. Choi,
Assessment of socioeconomic inequality based on virus-contaminated water usage in
developing countries: A review, Environ. Res. 192 (2021) 110309.
https://doi.org/https://doi.org/10.1016/j.envres.2020.110309.
[12] S. Lahrich, F. Laghrib, A. Farahi, M. Bakasse, S. Saqrane, M.A. El Mhammedi, Review
on the contamination of wastewater by COVID-19 virus: Impact and treatment, Sci. Total
Environ. 751 (2021) 142325.
https://doi.org/https://doi.org/10.1016/j.scitotenv.2020.142325.
[13] L. Pisharody, S. Suresh, S. Mukherji, Evaluation of adsorbents and eluents for application
in virus concentration and adsorption-desorption isotherms for coliphages, Chem. Eng. J.
403 (2021) 126267. https://doi.org/https://doi.org/10.1016/j.cej.2020.126267.
[14] C.P. Gerba, W.Q. Betancourt, M. Kitajima, How much reduction of virus is needed for
recycled water: A continuous changing need for assessment?, Water Res. 108 (2017) 25–
31. https://doi.org/https://doi.org/10.1016/j.watres.2016.11.020.
42
[15] A. Carducci, I. Federigi, D. Liu, J.R. Thompson, M. Verani, Making Waves: Coronavirus
detection, presence and persistence in the water environment: State of the art and
knowledge needs for public health, Water Res. 179 (2020) 115907.
https://doi.org/https://doi.org/10.1016/j.watres.2020.115907.
[16] D.H. Park, Y.H. Joe, A. Piri, S. An, J. Hwang, Determination of Air Filter Anti-Viral
Efficiency against an Airborne Infectious Virus, J. Hazard. Mater. 396 (2020) 122640.
https://doi.org/https://doi.org/10.1016/j.jhazmat.2020.122640.
[17] E. Haramoto, M. Kitajima, A. Hata, J.R. Torrey, Y. Masago, D. Sano, H. Katayama, A
review on recent progress in the detection methods and prevalence of human enteric
viruses in water, Water Res. 135 (2018) 168–186.
https://doi.org/https://doi.org/10.1016/j.watres.2018.02.004.
[18] C.M. Morrison, W.Q. Betancourt, D.R. Quintanar, G.U. Lopez, I.L. Pepper, C.P. Gerba,
Potential indicators of virus transport and removal during soil aquifer treatment of treated
wastewater effluent, Water Res. 177 (2020) 115812.
https://doi.org/https://doi.org/10.1016/j.watres.2020.115812.
[19] C.P. Gerba, W.Q. Betancourt, M. Kitajima, C.M. Rock, Reducing uncertainty in
estimating virus reduction by advanced water treatment processes, Water Res. 133 (2018)
282–288. https://doi.org/10.1016/j.watres.2018.01.044.
[20] S. Springthorpe, S.A. Sattar, Chapter 6 Virus Removal During Drinking Water Treatment,
Perspect. Med. Virol. 17 (2007) 109–126. https://doi.org/10.1016/S0168-7069(07)17006-
3.
[21] V. Nilsen, E. Christensen, M. Myrmel, A. Heistad, Spatio-temporal dynamics of virus and
bacteria removal in dual-media contact-filtration for drinking water, Water Res. 156
(2019) 9–22. https://doi.org/https://doi.org/10.1016/j.watres.2019.02.029.
[22] H.M.K. Delanka-Pedige, S.P. Munasinghe-Arachchige, Y. Zhang, N. Nirmalakhandan,
Bacteria and virus reduction in secondary treatment: Potential for minimizing post
disinfectant demand, Water Res. 177 (2020) 115802.
https://doi.org/https://doi.org/10.1016/j.watres.2020.115802.
[23] L. Casanova, W.A. Rutala, D.J. Weber, M.D. Sobsey, Survival of surrogate coronaviruses
in water, Water Res. 43 (2009) 1893–1898.
https://doi.org/https://doi.org/10.1016/j.watres.2009.02.002.
[24] S. Heider, C. Metzner, Quantitative real-time single particle analysis of virions, Virology.
462–463 (2014) 199–206. https://doi.org/10.1016/j.virol.2014.06.005.
[25] B. Michen, F. Meder, A. Rust, J. Fritsch, C. Aneziris, T. Graule, Virus removal in ceramic
43
depth filters based on diatomaceous earth, Environ. Sci. Technol. 46 (2012) 1173–1177.
https://doi.org/10.1021/es2030565.
[26] B. Michen, J. Fritsch, C. Aneziris, T. Graule, Improved virus removal in ceramic depth
filters modified with MgO, Environ. Sci. Technol. 47 (2013) 1526–1533.
https://doi.org/10.1021/es303685a.
[27] K.P. Goswami, G. Pugazhenthi, Credibility of polymeric and ceramic membrane filtration
in the removal of bacteria and virus from water: A review, J. Environ. Manage. 268
(2020) 110583. https://doi.org/10.1016/j.jenvman.2020.110583.
[28] N. Shirasaki, T. Matsushita, Y. Matsui, K. Murai, A. Aochi, Elimination of representative
contaminant candidate list viruses, coxsackievirus, echovirus, hepatitis A virus, and
norovirus, from water by coagulation processes, J. Hazard. Mater. 326 (2017) 110–119.
https://doi.org/https://doi.org/10.1016/j.jhazmat.2016.11.005.
[29] W.Q. Betancourt, J. Schijven, J. Regnery, A. Wing, C.M. Morrison, J.E. Drewes, C.P.
Gerba, Variable non-linear removal of viruses during transport through a saturated soil
column, J. Contam. Hydrol. 223 (2019) 103479.
https://doi.org/10.1016/j.jconhyd.2019.04.002.
[30] T. Prado, A. de Castro Bruni, M.R.F. Barbosa, S.C. Garcia, A.M. de Jesus Melo, M.I.Z.
Sato, Performance of wastewater reclamation systems in enteric virus removal, Sci. Total
Environ. 678 (2019) 33–42. https://doi.org/10.1016/j.scitotenv.2019.04.435.
[31] J.P.S. Sidhu, K. Sena, L. Hodgers, A. Palmer, S. Toze, Comparative enteric viruses and
coliphage removal during wastewater treatment processes in a sub-tropical environment,
Sci. Total Environ. 616–617 (2018) 669–677.
https://doi.org/10.1016/j.scitotenv.2017.10.265.
[32] W. Randazzo, P. Truchado, E. Cuevas-Ferrando, P. Simón, A. Allende, G. Sánchez,
SARS-CoV-2 RNA in wastewater anticipated COVID-19 occurrence in a low prevalence
area, Water Res. 181 (2020) 115942.
https://doi.org/https://doi.org/10.1016/j.watres.2020.115942.
[33] T. Miura, S. Okabe, Y. Nakahara, D. Sano, Removal properties of human enteric viruses
in a pilot-scale membrane bioreactor (MBR) process, Water Res. 75 (2015) 282–291.
https://doi.org/10.1016/j.watres.2015.02.046.
[34] C. Zhang, Y. Li, D. Shuai, Y. Shen, D. Wang, Progress and challenges in photocatalytic
disinfection of waterborne Viruses: A review to fill current knowledge gaps, Chem. Eng.
J. 355 (2019) 399–415. https://doi.org/https://doi.org/10.1016/j.cej.2018.08.158.
[35] A. Habibi-Yangjeh, S. Asadzadeh-Khaneghah, S. Feizpoor, A. Rouhi, Review on
44
heterogeneous photocatalytic disinfection of waterborne, airborne, and foodborne viruses:
Can we win against pathogenic viruses?, J. Colloid Interface Sci. 580 (2020) 503–514.
https://doi.org/10.1016/j.jcis.2020.07.047.
[36] G. La Rosa, L. Bonadonna, L. Lucentini, S. Kenmoe, E. Suffredini, Coronavirus in water
environments: Occurrence, persistence and concentration methods - A scoping review,
Water Res. 179 (2020) 115899.
https://doi.org/https://doi.org/10.1016/j.watres.2020.115899.
[37] E. Cheek, V. Guercio, C. Shrubsole, S. Dimitroulopoulou, Portable air purification:
Review of impacts on indoor air quality and health, Sci. Total Environ. (2020) 142585.
https://doi.org/10.1016/j.scitotenv.2020.142585.
[38] B. Wu, R. Wang, A.G. Fane, The roles of bacteriophages in membrane-based water and
wastewater treatment processes: A review, Water Res. 110 (2017) 120–132.
https://doi.org/10.1016/j.watres.2016.12.004.
[39] G. Carvajal, A. Branch, S.A. Sisson, D.J. Roser, B. van den Akker, P. Monis, P. Reeve, A.
Keegan, R. Regel, S.J. Khan, Virus removal by ultrafiltration: Understanding long-term
performance change by application of Bayesian analysis, Water Res. 122 (2017) 269–279.
https://doi.org/10.1016/j.watres.2017.05.057.
[40] R. Lu, C. Zhang, M. Piatkovsky, M. Ulbricht, M. Herzberg, T.H. Nguyen, Improvement
of virus removal using ultrafiltration membranes modified with grafted zwitterionic
polymer hydrogels, Water Res. 116 (2017) 86–94.
https://doi.org/10.1016/j.watres.2017.03.023.
[41] G.A. Junter, L. Lebrun, Cellulose-based virus-retentive filters: a review, Rev. Environ.
Sci. Biotechnol. 16 (2017) 455–489. https://doi.org/10.1007/s11157-017-9434-1.
[42] M. Wegmann, B. Michen, T. Luxbacher, J. Fritsch, T. Graule, Modification of ceramic
microfilters with colloidal zirconia to promote the adsorption of viruses from water, Water
Res. 42 (2008) 1726–1734. https://doi.org/10.1016/j.watres.2007.10.030.
[43] S. Bofill-Mas, M. Rusiñol, Recent trends on methods for the concentration of viruses from
water samples, Curr. Opin. Environ. Sci. Heal. (2020).
https://doi.org/10.1016/j.coesh.2020.01.006.
[44] L. Manukyan, J. Padova, A. Mihranyan, Virus removal filtration of chemically defined
Chinese Hamster Ovary cells medium with nanocellulose-based size exclusion filter,
Biologicals. 59 (2019) 62–67. https://doi.org/10.1016/j.biologicals.2019.03.001.
[45] F. Rambags, C.C. Tanner, R. Stott, L.A. Schipper, Bacteria and virus removal in
denitrifying bioreactors: Effects of media type and age, Ecol. Eng. 138 (2019) 46–53.
45
https://doi.org/10.1016/j.ecoleng.2019.06.021.
[46] Q.L. Shimabuku, T. Ueda-Nakamura, R. Bergamasco, M.R. Fagundes-Klen, Chick-
Watson kinetics of virus inactivation with granular activated carbon modified with silver
nanoparticles and/or copper oxide, Process Saf. Environ. Prot. 117 (2018) 33–42.
https://doi.org/10.1016/j.psep.2018.04.005.
[47] T. Matsushita, H. Suzuki, N. Shirasaki, Y. Matsui, K. Ohno, Adsorptive virus removal
with super-powdered activated carbon, Sep. Purif. Technol. 107 (2013) 79–84.
https://doi.org/10.1016/j.seppur.2013.01.017.
[48] T. Powell, G.M. Brion, M. Jagtoyen, F. Derbyshire, Investigating the effect of carbon
shape on virus adsorption, Environ. Sci. Technol. 34 (2000) 2779–2783.
https://doi.org/10.1021/es991097w.
[49] Q.L. Shimabuku, F.S. Arakawa, M. Fernandes Silva, P. Ferri Coldebella, T. Ueda-
Nakamura, M.R. Fagundes-Klen, R. Bergamasco, Water treatment with exceptional virus
inactivation using activated carbon modified with silver (Ag) and copper oxide (CuO)
nanoparticles, Environ. Technol. (United Kingdom). 38 (2017) 2058–2069.
https://doi.org/10.1080/09593330.2016.1245361.
[50] G.A. Junter, L. Lebrun, Polysaccharide-based chromatographic adsorbents for virus
purification and viral clearance, J. Pharm. Anal. 10 (2020) 291–312.
https://doi.org/10.1016/j.jpha.2020.01.002.
[51] Y. Xing, A. Ellis, M. Magnuson, W.F. Harper, Adsorption of bacteriophage MS2 to
colloids: Kinetics and particle interactions, Colloids Surfaces A Physicochem. Eng. Asp.
585 (2020) 124099. https://doi.org/10.1016/j.colsurfa.2019.124099.
[52] Y.H. Joe, D.H. Park, J. Hwang, Evaluation of Ag nanoparticle coated air filter against
aerosolized virus: Anti-viral efficiency with dust loading, J. Hazard. Mater. 301 (2016)
547–553. https://doi.org/https://doi.org/10.1016/j.jhazmat.2015.09.017.
[53] M. Minoshima, Y. Lu, T. Kimura, R. Nakano, H. Ishiguro, Y. Kubota, K. Hashimoto, K.
Sunada, Comparison of the antiviral effect of solid-state copper and silver compounds, J.
Hazard. Mater. 312 (2016) 1–7.
https://doi.org/https://doi.org/10.1016/j.jhazmat.2016.03.023.
[54] C. V Chrysikopoulos, A.F. Aravantinou, Virus inactivation in the presence of quartz sand
under static and dynamic batch conditions at different temperatures, J. Hazard. Mater.
233–234 (2012) 148–157. https://doi.org/https://doi.org/10.1016/j.jhazmat.2012.07.002.
[55] X. Mi, C.L. Heldt, Adsorption of a non-enveloped mammalian virus to functionalized
nanofibers, Colloids Surfaces B Biointerfaces. 121 (2014) 319–324.
46
https://doi.org/10.1016/j.colsurfb.2014.06.007.
[56] J.W. Chung, J.W. Foppen, G. Gerner, R. Krebs, P.N.L. Lens, Removal of rotavirus and
adenovirus from artificial ground water using hydrochar derived from sewage sludge, J.
Appl. Microbiol. 119 (2015) 876–884. https://doi.org/10.1111/jam.12863.
[57] X. Zheng, D. Chen, zhiwei Wang, Y. Lei, R. Cheng, Nano-TiO2 membrane adsorption
reactor (MAR) for virus removal in drinking water, Chem. Eng. J. 230 (2013) 180–187.
https://doi.org/10.1016/j.cej.2013.06.069.
[58] G. Tiliket, D. Le Sage, V. Moules, M. Rosa-Calatrava, B. Lina, J.M. Valleton, Q.T.
Nguyen, L. Lebrun, A new material for airborne virus filtration, Chem. Eng. J. 173 (2011)
341–351. https://doi.org/https://doi.org/10.1016/j.cej.2011.07.059.
[59] R. Cheng, M. Kang, S. Zhuang, S. Wang, X. Zheng, X. Pan, L. Shi, J. Wang, Removal of
bacteriophage f2 in water by Fe/Ni nanoparticles: Optimization of Fe/Ni ratio and
influencing factors, Sci. Total Environ. 649 (2019) 995–1003.
https://doi.org/10.1016/j.scitotenv.2018.08.380.
[60] K.R. Bright, E.E. Sicairos-Ruelas, P.M. Gundy, C.P. Gerba, Assessment of the Antiviral
Properties of Zeolites Containing Metal Ions, Food Environ. Virol. 1 (2009) 37–41.
https://doi.org/10.1007/s12560-008-9006-1.
[61] P.J. Chisholm, J.W. Busch, D.W. Crowder, Effects of life history and ecology on virus
evolutionary potential, Virus Res. 265 (2019) 1–9.
https://doi.org/10.1016/j.virusres.2019.02.018.
[62] R. Zandi, B. Dragnea, A. Travesset, R. Podgornik, On virus growth and form, Phys. Rep.
847 (2020) 1–102. https://doi.org/10.1016/j.physrep.2019.12.005.
[63] P. Gamazo, M. Victoria, J.F. Schijven, E. Alvareda, L.F.L. Tort, J. Ramos, L.A.
Lizasoain, G. Sapriza, M. Castells, L. Bessone, R. Colina, Modeling the Transport of
Human Rotavirus and Norovirus in Standardized and in Natural Soil Matrix-Water
Systems, Food Environ. Virol. 12 (2020) 58–67. https://doi.org/10.1007/s12560-019-
09414-z.
[64] L. Manukyan, P. Li, S. Gustafsson, A. Mihranyan, Growth media filtration using
nanocellulose-based virus removal filter for upstream biopharmaceutical processing, J.
Memb. Sci. 572 (2019) 464–474. https://doi.org/10.1016/j.memsci.2018.11.002.
[65] J. Bartels, A.G. Batista, S. Kroll, M. Maas, K. Rezwan, Hydrophobic ceramic capillary
membranes for versatile virus filtration, J. Memb. Sci. 570–571 (2019) 85–92.
https://doi.org/10.1016/j.memsci.2018.10.022.
[66] A. Armanious, M. Aeppli, R. Jacak, D. Refardt, T. Sigstam, T. Kohn, M. Sander, Viruses
47
at Solid-Water Interfaces: A Systematic Assessment of Interactions Driving Adsorption,
Environ. Sci. Technol. 50 (2016) 732–743. https://doi.org/10.1021/acs.est.5b04644.
[67] F. Meder, J. Wehling, A. Fink, B. Piel, K. Li, K. Frank, A. Rosenauer, L. Treccani, S.
Koeppen, A. Dotzauer, K. Rezwan, The role of surface functionalization of colloidal
alumina particles on their controlled interactions with viruses, Biomaterials. 34 (2013)
4203–4213. https://doi.org/10.1016/j.biomaterials.2013.02.059.
[68] C.L. Heldt, A. Zahid, K.S. Vijayaragavan, X. Mi, Experimental and computational surface
hydrophobicity analysis of a non-enveloped virus and proteins, Colloids Surfaces B
Biointerfaces. 153 (2017) 77–84. https://doi.org/10.1016/j.colsurfb.2017.02.011.
[69] X. Mi, E.K. Bromley, P.U. Joshi, F. Long, C.L. Heldt, Virus Isoelectric Point
Determination Using Single-Particle Chemical Force Microscopy, Langmuir. 36 (2020)
370–378. https://doi.org/10.1021/acs.langmuir.9b03070.
[70] M.E. Verbyla, J.R. Mihelcic, A review of virus removal in wastewater treatment pond
systems, Water Res. 71 (2015) 107–124. https://doi.org/10.1016/j.watres.2014.12.031.
[71] Z. Németh, G.P. Szekeres, M. Schabikowski, K. Schrantz, J. Traber, W. Pronk, K.
Hernádi, T. Graule, Enhanced virus filtration in hybrid membranes with MWCNT
nanocomposite, R. Soc. Open Sci. 6 (2019) 181294. doi:10.1098/rsos.181294.
[72] L. Gutierrez, X. Li, J. Wang, G. Nangmenyi, J. Economy, T.B. Kuhlenschmidt, M.S.
Kuhlenschmidt, T.H. Nguyen, Adsorption of rotavirus and bacteriophage MS2 using glass
fiber coated with hematite nanoparticles, Water Res. 43 (2009) 5198–5208.
https://doi.org/10.1016/j.watres.2009.08.031.
[73] J.M. Mazurkow, N.S. Yüzbasi, K.W. Domagala, S. Pfeiffer, D. Kata, T. Graule, Nano-
Sized Copper (Oxide) on Alumina Granules for Water Filtration: Effect of Copper
Oxidation State on Virus Removal Performance, Environ. Sci. Technol. 54 (2020) 1214–
1222. https://doi.org/10.1021/acs.est.9b05211.
[74] I. Bradley, A. Straub, P. Maraccini, S. Markazi, T.H. Nguyen, Iron oxide amended
biosand filters for virus removal, Water Res. 45 (2011) 4501–4510.
https://doi.org/10.1016/j.watres.2011.05.045.
[75] K. Domagała, C. Jacquin, M. Borlaf, B. Sinnet, T. Julian, D. Kata, T. Graule, Efficiency
and stability evaluation of Cu2O/MWCNTs filters for virus removal from water, Water
Res. 179 (2020) 115879. https://doi.org/https://doi.org/10.1016/j.watres.2020.115879.
[76] M. V. Liga, S.J. Maguire-Boyle, H.R. Jafry, A.R. Barron, Q. Li, Silica decorated TiO2 for
virus inactivation in drinking water -simple synthesis method and mechanisms of
enhanced inactivation kinetics, Environ. Sci. Technol. 47 (2013) 6463–6470.
48
https://doi.org/10.1021/es400196p.
[77] S. Mintova, M. Jaber, V. Valtchev, Nanosized microporous crystals: emerging
applications, Chem. Soc. Rev. 44 (2015) 7207–7233. https://doi.org/10.1039/c5cs00210a.
[78] H. Awala, J.P. Gilson, R. Retoux, P. Boullay, J.M. Goupil, V. Valtchev, S. Mintova,
Template-free nanosized faujasite-type zeolites, Nat. Mater. 14 (2015) 447–451.
https://doi.org/10.1038/nmat4173.
[79] E.P. Ng, D. Chateigner, T. Bein, V. Valtchev, S. Mintova, Capturing ultrasmall EMT
zeolite from template-free systems, Science. 335 (2012) 70–73.
https://doi.org/10.1126/science.1214798.
[80] V. Georgieva, C. Anfray, R. Retoux, V. Valtchev, S. Valable, S. Mintova, Iron loaded
EMT nanosized zeolite with high affinity towards CO2 and NO, Microporous
Mesoporous Mater. 232 (2016) 256–263.
https://doi.org/10.1016/J.MICROMESO.2016.06.015.
[81] P. Losch, H.R. Joshi, O. Vozniuk, A. Grünert, C. Ochoa-Hernández, H. Jabraoui, M.
Badawi, W. Schmidt, Proton Mobility, Intrinsic Acid Strength, and Acid Site Location in
Zeolites Revealed by Varying Temperature Infrared Spectroscopy and Density Functional
Theory Studies, J. Am. Chem. Soc. 140 (2018) 17790–17799.
https://doi.org/10.1021/jacs.8b11588.
[82] S. Chibani, I. Medlej, S. Lebègue, J.G. Ángyán, L. Cantrel, M. Badawi, Performance of
CuII-, PbII-, and HgII-Exchanged Mordenite in the Adsorption of I2, ICH3, H2O, CO,
ClCH3, and Cl2: A Density Functional Study, ChemPhysChem. 18 (2017) 1642–1652.
https://doi.org/10.1002/cphc.201700104.
[83] E.P. Hessou, W.G. Kanhounnon, D. Rocca, H. Monnier, C. Vallières, S. Lebègue, M.
Badawi, Adsorption of NO, NO2, CO, H2O and CO2 over isolated monovalent cations in
faujasite zeolite: a periodic DFT investigation, Theor. Chem. Acc. 137 (2018) 161.
https://doi.org/10.1007/s00214-018-2373-2.
[84] H. Jabraoui, I. Khalil, S. Lebègue, M. Badawi, Ab initio screening of cation-exchanged
zeolites for biofuel purification, Mol. Syst. Des. Eng. 4 (2019) 882–892.
https://doi.org/10.1039/c9me00015a.
[85] I. Khalil, H. Jabraoui, S. Lebègue, W.J. Kim, L.J. Aguilera, K. Thomas, F. Maugé, M.
Badawi, Biofuel purification: Coupling experimental and theoretical investigations for
efficient separation of phenol from aromatics by zeolites, Chem. Eng. J. 402 (2020)
126264. https://doi.org/10.1016/j.cej.2020.126264.
[86] R. Marega, E.A. Prasetyanto, C. Michiels, L. De Cola, D. Bonifazi, Fast Targeting and
49
Cancer Cell Uptake of Luminescent, (2016) 1–11.
https://doi.org/10.1002/smll.201601447.
[87] L. Pang, K. Farkas, G. Bennett, A. Varsani, R. Easingwood, R. Tilley, U. Nowostawska,
S. Lin, Mimicking filtration and transport of rotavirus and adenovirus in sand media using
DNA-labeled, protein-coated silica nanoparticles, Water Res. 62 (2014) 167–179.
https://doi.org/https://doi.org/10.1016/j.watres.2014.05.055.
[88] G. Marinescu, D.C. Culita, C. Romanitan, S. Somacescu, C.D. Ene, V. Marinescu, D.G.
Negreanu, C. Maxim, M. Popa, L. Marutescu, M. Stan, C. Chifiriuc, Novel hybrid
materials based on heteroleptic Ru(III) complexes immobilized on SBA-15 mesoporous
silica as highly potent antimicrobial and cytotoxic agents, Appl. Surf. Sci. 520 (2020)
146379. https://doi.org/https://doi.org/10.1016/j.apsusc.2020.146379.
[89] Y. Qin, Z. Wen, W. Zhang, J. Chai, D. Liu, S. Wu, Different roles of silica nanoparticles
played in virus transport in saturated and unsaturated porous media, Environ. Pollut. 259
(2020) 113861. https://doi.org/https://doi.org/10.1016/j.envpol.2019.113861.
[90] S.A. Jadhav, H.B. Garud, A.H. Patil, G.D. Patil, C.R. Patil, T.D. Dongale, P.S. Patil,
Recent advancements in silica nanoparticles based technologies for removal of dyes from
water, Colloid Interface Sci. Commun. 30 (2019) 100181.
https://doi.org/https://doi.org/10.1016/j.colcom.2019.100181.
[91] P.N.E. Diagboya, E.D. Dikio, Silica-based mesoporous materials; emerging designer
adsorbents for aqueous pollutants removal and water treatment, Microporous Mesoporous
Mater. 266 (2018) 252–267.
https://doi.org/https://doi.org/10.1016/j.micromeso.2018.03.008.
[92] M.M. Abeer, P. Rewatkar, Z. Qu, M. Talekar, F. Kleitz, R. Schmid, M. Lindén, T.
Kumeria, A. Popat, Silica nanoparticles: A promising platform for enhanced oral delivery
of macromolecules, J. Control. Release. 326 (2020) 544–555.
https://doi.org/https://doi.org/10.1016/j.jconrel.2020.07.021.
[93] T. Poeckh, S. Lopez, A.O. Fuller, M.J. Solomon, R.G. Larson, Adsorption and elution
characteristics of nucleic acids on silica surfaces and their use in designing a miniaturized
purification unit, Anal. Biochem. 373 (2008) 253–262.
https://doi.org/https://doi.org/10.1016/j.ab.2007.10.026.
[94] H. Clemens, L. Pang, L.K. Morgan, L. Weaver, Attenuation of rotavirus, MS2
bacteriophage and biomolecule-modified silica nanoparticles in undisturbed silt loam over
gravels dosed with onsite wastewater, Water Res. 169 (2020) 115272.
https://doi.org/https://doi.org/10.1016/j.watres.2019.115272.
50
[95] M. Liu, A. Hata, H. Katayama, I. Kasuga, Consecutive ultrafiltration and silica adsorption
for recovery of extracellular antibiotic resistance genes from an urban river, Environ.
Pollut. 260 (2020) 114062. https://doi.org/https://doi.org/10.1016/j.envpol.2020.114062.
[96] I. Berts, G. Fragneto, L. Porcar, M.S. Hellsing, A.R. Rennie, Controlling adsorption of
albumin with hyaluronan on silica surfaces and sulfonated latex particles, J. Colloid
Interface Sci. 504 (2017) 315–324.
https://doi.org/https://doi.org/10.1016/j.jcis.2017.05.037.
[97] S. Bone, A. Alum, J. Markovski, K. Hristovski, E. Bar-Zeev, Y. Kaufman, M.
Abbaszadegan, F. Perreault, Physisorption and chemisorption of T4 bacteriophages on
amino functionalized silica particles, J. Colloid Interface Sci. 532 (2018) 68–76.
https://doi.org/https://doi.org/10.1016/j.jcis.2018.07.107.
[98] N. Sun, C. Deng, Y. Liu, X. Zhao, Y. Tang, R. Liu, Q. Xia, W. Yan, G. Ge, Optimization
of influencing factors of nucleic acid adsorption onto silica-coated magnetic particles:
Application to viral nucleic acid extraction from serum, J. Chromatogr. A. 1325 (2014)
31–39. https://doi.org/https://doi.org/10.1016/j.chroma.2013.11.059.
[99] M. Sun, H. Wang, X. Li, Modification of cellulose microfibers by polyglutamic acid and
mesoporous silica nanoparticles for Enterovirus 71 adsorption, Mater. Lett. 277 (2020)
128320. https://doi.org/https://doi.org/10.1016/j.matlet.2020.128320.
[100] J. Kim, H. Lee, J.-Y. Lee, K.-H. Park, W. Kim, J.H. Lee, H.-J. Kang, S.W. Hong, H.-J.
Park, S. Lee, J.-H. Lee, H.-D. Park, J.Y. Kim, Y.W. Jeong, J. Lee, Photosensitized
Production of Singlet Oxygen via C60 Fullerene Covalently Attached to Functionalized
Silica-coated Stainless-Steel Mesh: Remote Bacterial and Viral Inactivation, Appl. Catal.
B Environ. 270 (2020) 118862.
https://doi.org/https://doi.org/10.1016/j.apcatb.2020.118862.
[101] F.E. Galdino, A.S. Picco, M.L. Sforca, M.B. Cardoso, W. Loh, Effect of particle
functionalization and solution properties on the adsorption of bovine serum albumin and
lysozyme onto silica nanoparticles, Colloids Surfaces B Biointerfaces. 186 (2020) 110677.
https://doi.org/https://doi.org/10.1016/j.colsurfb.2019.110677.
[102] S. Ghosh, V. Vandana, Nano-structured mesoporous silica/silver composite: Synthesis,
characterization and targeted application towards water purification, Mater. Res. Bull. 88
(2017) 291–300. https://doi.org/https://doi.org/10.1016/j.materresbull.2016.12.044.
[103] A. Kurtz-Chalot, C. Villiers, J. Pourchez, D. Boudard, M. Martini, P.N. Marche, M.
Cottier, V. Forest, Impact of silica nanoparticle surface chemistry on protein corona
formation and consequential interactions with biological cells, Mater. Sci. Eng. C. 75
51
(2017) 16–24. https://doi.org/https://doi.org/10.1016/j.msec.2017.02.028.
[104] Z. Chen, F.-C. Hsu, D. Battigelli, H.-C. Chang, Capture and release of viruses using
amino-functionalized silica particles, Anal. Chim. Acta. 569 (2006) 76–82.
https://doi.org/https://doi.org/10.1016/j.aca.2006.03.103.
[105] K. He, C. Chen, C. Liang, C. Liu, B. Yang, X. Chen, C. Cai, Highly selective recognition
and fluorescent detection of JEV via virus-imprinted magnetic silicon microspheres,
Sensors Actuators B Chem. 233 (2016) 607–614.
https://doi.org/https://doi.org/10.1016/j.snb.2016.04.127.
[106] S. Zhan, Y. Yang, Z. Shen, J. Shan, Y. Li, S. Yang, D. Zhu, Efficient removal of
pathogenic bacteria and viruses by multifunctional amine-modified magnetic
nanoparticles, J. Hazard. Mater. 274 (2014) 115–123.
https://doi.org/https://doi.org/10.1016/j.jhazmat.2014.03.067.
[107] R. Cademartiri, H. Anany, I. Gross, R. Bhayani, M. Griffiths, M.A. Brook,
Immobilization of bacteriophages on modified silica particles, Biomaterials. 31 (2010)
1904–1910. https://doi.org/https://doi.org/10.1016/j.biomaterials.2009.11.029.
[108] R. Riccò, W. Liang, S. Li, J.J. Gassensmith, F. Caruso, C. Doonan, P. Falcaro, Metal-
Organic Frameworks for Cell and Virus Biology: A Perspective, ACS Nano. 12 (2018)
13–23. https://doi.org/10.1021/acsnano.7b08056.
[109] C. Doonan, R. Riccò, K. Liang, D. Bradshaw, P. Falcaro, Metal-Organic Frameworks at
the Biointerface: Synthetic Strategies and Applications, Acc. Chem. Res. 50 (2017) 1423–
1432. https://doi.org/10.1021/acs.accounts.7b00090.
[110] K. Adil, Y. Belmabkhout, R.S. Pillai, A. Cadiau, P.M. Bhatt, A.H. Assen, G. Maurin, M.
Eddaoudi, Gas/vapour separation using ultra-microporous metal-organic frameworks:
Insights into the structure/separation relationship, Chem. Soc. Rev. 46 (2017) 3402–3430.
https://doi.org/10.1039/c7cs00153c.
[111] S. Chibani, M. Badawi, T. Loiseau, C. Volkringer, L. Cantrel, J.F. Paul, A DFT study of
RuO4 interactions with porous materials: Metal-organic frameworks (MOFs) and zeolites,
Phys. Chem. Chem. Phys. 20 (2018) 16770–16776. https://doi.org/10.1039/c8cp01950a.
[112] K. Liang, J.J. Richardson, J. Cui, F. Caruso, C.J. Doonan, P. Falcaro, Biomimetics:
Metal–Organic Framework Coatings as Cytoprotective Exoskeletons for Living Cells,
Adv. Mater. 28 (2016) 8066. https://doi.org/10.1002/adma.201670256.
[113] B.-P. Xie, G.-H. Qiu, P.-P. Hu, Z. Liang, Y.-M. Liang, B. Sun, L.-P. Bai, Z.-H. Jiang, J.-
X. Chen, Simultaneous detection of Dengue and Zika virus RNA sequences with a three-
dimensional Cu-based zwitterionic metal–organic framework, comparison of single and
52
synchronous fluorescence analysis, Sensors Actuators B Chem. 254 (2018) 1133–1140.
https://doi.org/https://doi.org/10.1016/j.snb.2017.06.085.
[114] Y.-W. Zhang, W.-S. Liu, J.-S. Chen, H.-L. Niu, C.-J. Mao, B.-K. Jin, Metal-organic gel
and metal-organic framework based switchable electrochemiluminescence RNA sensing
platform for Zika virus, Sensors Actuators B Chem. 321 (2020) 128456.
https://doi.org/https://doi.org/10.1016/j.snb.2020.128456.
[115] Z. Jia, Y. Ma, L. Yang, C. Guo, N. Zhou, M. Wang, L. He, Z. Zhang, NiCo2O4 spinel
embedded with carbon nanotubes derived from bimetallic NiCo metal-organic framework
for the ultrasensitive detection of human immune deficiency virus-1 gene, Biosens.
Bioelectron. 133 (2019) 55–63. https://doi.org/https://doi.org/10.1016/j.bios.2019.03.030.
[116] J. Yang, W. Feng, K. Liang, C. Chen, C. Cai, A novel fluorescence molecularly imprinted
sensor for Japanese encephalitis virus detection based on metal organic frameworks and
passivation-enhanced selectivity, Talanta. 212 (2020) 120744.
https://doi.org/https://doi.org/10.1016/j.talanta.2020.120744.
[117] K.J. Clark, A.B. Sarr, P.G. Grant, T.D. Phillips, G.N. Woode, In vitro studies on the use of
clay, clay minerals and charcoal to adsorb bovine rotavirus and bovine coronavirus, Vet.
Microbiol. 63 (1998) 137–146. https://doi.org/10.1016/S0378-1135(98)00241-7.
[118] M. Knappe, S. Bodevin, H.C. Selinka, D. Spillmann, R.E. Streeck, X.S. Chen, U. Lindahl,
M. Sapp, Surface-exposed amino acid residues of HPV16 L1 protein mediating interaction
with cell surface heparan sulfate, J. Biol. Chem. 282 (2007) 27913–27922.
https://doi.org/10.1074/jbc.M705127200.
[119] V. Cagno, P. Andreozzi, M. D’Alicarnasso, P.J. Silva, M. Mueller, M. Galloux, R. Le
Goffic, S.T. Jones, M. Vallino, J. Hodek, J. Weber, S. Sen, E.R. Janecek, A. Bekdemir, B.
Sanavio, C. Martinelli, M. Donalisio, M.A.R. Welti, J.F. Eleouet, Y. Han, L. Kaiser, L.
Vukovic, C. Tapparel, P. Král, S. Krol, D. Lembo, F. Stellacci, Broad-spectrum non-toxic
antiviral nanoparticles with a virucidal inhibition mechanism, Nat. Mater. 17 (2018) 195–
203. https://doi.org/10.1038/NMAT5053.
[120] S.T. Jones, V. Cagno, M. Janeček, D. Ortiz, N. Gasilova, J. Piret, M. Gasbarri, D.A.
Constant, Y. Han, L. Vuković, P. Král, L. Kaiser, S. Huang, S. Constant, K. Kirkegaard,
G. Boivin, F. Stellacci, C. Tapparel, Modified cyclodextrins as broad-spectrum antivirals,
Sci. Adv. 6 (2020) eaax9318. https://doi.org/10.1126/sciadv.aax9318.
[121] S.K. Kim, Y.X. Li, Medicinal benefits of sulfated polysaccharides from sea vegetables,
Adv. Food Nutr. Res. 64 (2011) 391–402. https://doi.org/10.1016/B978-0-12-387669-
0.00030-2.
53
[122] L.M. Harhaji Trajkovic, S.A. Mijatovic, D.D. Maksimovic-Ivanic, I.D. Stojanovic, M.B.
Momcilovic, S.J. Tufegdzic, V.M. Maksimovic, Z.S. Marjanovi, S.D. Stosic-Grujicic,
Anticancer properties of ganoderma lucidum methanol extracts in vitro and in vivo, Nutr.
Cancer. 61 (2009) 696–707. https://doi.org/10.1080/01635580902898743.
[123] Y. Xu, Q. Dong, H. Qiu, R. Cong, K. Ding, Structural characterization of an
arabinogalactan from Platycodon grandiflorum roots and antiangiogenic activity of its
sulfated derivative, Biomacromolecules. 11 (2010) 2558–2566.
https://doi.org/10.1021/bm100402n.
[124] L. Wang, X. Wang, H. Wu, R. Liu, Overview on biological activities and molecular
characteristics of sulfated polysaccharides from marine green algae in recent years, Mar.
Drugs. 12 (2014) 4984–5020. doi:10.3390/md12094984.
[125] M.H. Peng, W.P. Dai, S.J. Liu, L.W. Yu, Y.N. Wu, R. Liu, X.L. Chen, X.P. Lai, X. Li,
Z.X. Zhao, G. Li, Bioactive glycosides from the roots of Ilex asprella, Pharm. Biol. 54
(2016) 2127–2134. https://doi.org/10.3109/13880209.2016.1146779.
[126] M. Akram, I.M. Tahir, S.M.A. Shah, Z. Mahmood, A. Altaf, K. Ahmad, N. Munir, M.
Daniyal, S. Nasir, H. Mehboob, Antiviral potential of medicinal plants against HIV, HSV,
influenza, hepatitis, and coxsackievirus: A systematic review, Phyther. Res. 32 (2018)
811–822. https://doi.org/10.1002/ptr.6024.
[127] A.K. Singh, A.S. Bhadauria, P. Kumar, H. Bera, S. Saha, Bioactive and drug-delivery
potentials of polysaccharides and their derivatives, in: S. Maiti, S. Jana (Eds.), Polysacch.
Carriers Drug Deliv., Woodhead Publishing, 2019: pp. 19–48. doi:10.1016/B978-0-08-
102553-6.00002-7.
[128] W.A.J.P. Wijesinghe, Y.J. Jeon, Biological activities and potential industrial applications
of fucose rich sulfated polysaccharides and fucoidans isolated from brown seaweeds: A
review, Carbohydr. Polym. 88 (2012) 13–20.
https://doi.org/10.1016/j.carbpol.2011.12.029.
[129] V.D. Prajapati, P.M. Maheriya, G.K. Jani, H.K. Solanki, Carrageenan: A natural seaweed
polysaccharide and its applications, Carbohydr. Polym. 105 (2014) 97–112.
https://doi.org/10.1016/j.carbpol.2014.01.067.
[130] Y. Seki, M. Mizukura, T. Ichimiya, Y. Suda, S. Nishihara, M. Masuda, S. Takase-Yoden,
O-sulfate groups of heparin are critical for inhibition of ecotropic murine leukemia virus
infection by heparin, Virology. 424 (2012) 56–66.
https://doi.org/10.1016/j.virol.2011.11.030.
[131] J. Liu, S. Willför, C. Xu, A review of bioactive plant polysaccharides: Biological
54
activities, functionalization, and biomedical applications, Bioact. Carbohydrates Diet.
Fibre. 5 (2015) 31–61. https://doi.org/10.1016/j.bcdf.2014.12.001.
[132] L.C. Faccin-Galhardi, K. Aimi Yamamoto, S. Ray, B. Ray, R.E. Carvalho Linhares, C.
Nozawa, The in vitro antiviral property of Azadirachta indica polysaccharides for
poliovirus, J. Ethnopharmacol. 142 (2012) 86–90.
https://doi.org/10.1016/j.jep.2012.04.018.
[133] F. Ma, S. Kong, H. Tan, R. Wu, B. Xia, Y. Zhou, H. Xu, Structural characterization and
antiviral effect of a novel polysaccharide PSP-2B from Prunellae Spica, Carbohydr.
Polym. 152 (2016) 699–709. doi:10.1016/j.carbpol.2016.07.062.
[134] J.T. Richards, E.R. Kern, L.A. Glasgow, J.C. Overall, E.F. Deign, M.T. Hatch, Antiviral
activity of extracts from marine algae, Antimicrob. Agents Chemother. 14 (1978) 24–30.
https://doi.org/10.1128/AAC.14.1.24.
[135] M. Witvrouw, E. De Clercq, Sulfated polysaccharides extracted from sea algae as
potential antiviral drugs, Gen. Pharmacol. 29 (1997) 497–511.
https://doi.org/10.1016/S0306-3623(96)00563-0.
[136] L. Ono, W. Wollinger, I.M. Rocco, T.L.M. Coimbra, P.A.J. Gorin, M.R. Sierakowski, In
vitro and in vivo antiviral properties of sulfated galactomannans against yellow fever virus
(BeH111 strain) and dengue 1 virus (Hawaii strain), Antiviral Res. 60 (2003) 201–208.
https://doi.org/10.1016/S0166-3542(03)00175-X.
[137] K.I.P.J. Hidari, N. Takahashi, M. Arihara, M. Nagaoka, K. Morita, T. Suzuki, Structure
and anti-dengue virus activity of sulfated polysaccharide from a marine alga, Biochem.
Biophys. Res. Commun. 376 (2008) 91–95. https://doi.org/10.1016/j.bbrc.2008.08.100.
[138] P. Ghosh, U. Adhikari, P.K. Ghosal, C.A. Pujol, M.J. Carlucci, E.B. Damonte, B. Ray, In
vitro anti-herpetic activity of sulfated polysaccharide fractions from Caulerpa racemosa,
Phytochemistry. 65 (2004) 3151–3157. https://doi.org/10.1016/j.phytochem.2004.07.025.
[139] P. Mandal, C.A. Pujol, M.J. Carlucci, K. Chattopadhyay, E.B. Damonte, B. Ray, Anti-
herpetic activity of a sulfated xylomannan from Scinaia hatei, Phytochemistry. 69 (2008)
2193–2199. https://doi.org/10.1016/j.phytochem.2008.05.004.
[140] B.G. Malagoli, F.T.G.S. Cardozo, J.H.S. Gomes, V.P. Ferraz, C.M.O. Simões, F.C. Braga,
Chemical characterization and antiherpes activity of sulfated polysaccharides from
Lithothamnion muelleri, Int. J. Biol. Macromol. 66 (2014) 332–337.
https://doi.org/10.1016/j.ijbiomac.2014.02.053.
[141] T. Ghosh, K. Chattopadhyay, M. Marschall, P. Karmakar, P. Mandal, B. Ray, Focus on
antivirally active sulfated polysaccharides: From structure-activity analysis to clinical
55
evaluation, Glycobiology. 19 (2009) 2–15. https://doi.org/10.1093/glycob/cwn092.
[142] E. Heylen, J. Neyts, D. Jochmans, Drug candidates and model systems in respiratory
syncytial virus antiviral drug discovery, Biochem. Pharmacol. 127 (2017) 1–12.
https://doi.org/10.1016/j.bcp.2016.09.014.
[143] K. Dhama, K. Karthik, R. Khandia, A. Munjal, R. Tiwari, R. Rana, S.K. Khurana, S.
Ullah, R.U. Khan, M. Alagawany, M.R. Farag, M.D. and S.K. Joshi, Medicinal and
Therapeutic Potential of Herbs and Plant Metabolites / Extracts Countering Viral
Pathogens - Current Knowledge and Future Prospects, Curr. Drug Metab. 19 (2018) 236–
263. doi:http://dx.doi.org/10.2174/1389200219666180129145252.
[144] J.S. Yang, Y.J. Xie, W. He, Research progress on chemical modification of alginate: A
review, Carbohydr. Polym. 84 (2011) 33–39.
https://doi.org/10.1016/j.carbpol.2010.11.048.
[145] E. Damonte, M. Matulewicz, A. Cerezo, Sulfated Seaweed Polysaccharides as Antiviral
Agents, Curr. Med. Chem. 11 (2012) 2399–2419.
https://doi.org/10.2174/0929867043364504.
[146] U. Adhikari, C.G. Mateu, K. Chattopadhyay, C.A. Pujol, E.B. Damonte, B. Ray, Structure
and antiviral activity of sulfated fucans from Stoechospermum marginatum,
Phytochemistry. 67 (2006) 2474–2482. https://doi.org/10.1016/j.phytochem.2006.05.024.
[147] R. Copeland, A. Balasubramaniam, V. Tiwari, F. Zhang, A. Bridges, R.J. Linhardt, D.
Shukla, J. Liu, Using a 3-O-sulfated heparin octasaccharide to inhibit the entry of herpes
simplex virus type 1, Biochemistry. 47 (2008) 5774–5783.
https://doi.org/10.1021/bi800205t.
[148] S. Sinha, A. Astani, T. Ghosh, P. Schnitzler, B. Ray, Polysaccharides from Sargassum
tenerrimum: Structural features, chemical modification and anti-viral activity,
Phytochemistry. 71 (2010) 235–242. https://doi.org/10.1016/j.phytochem.2009.10.014.
[149] W. Tang, C. Zou, C. Da, Y. Cao, H. Peng, A review on the recent development of
cyclodextrin-based materials used in oilfield applications, Carbohydr. Polym. 240 (2020)
116321. https://doi.org/https://doi.org/10.1016/j.carbpol.2020.116321.
[150] M. Mahjoubin-Tehran, P.T. Kovanen, S. Xu, T. Jamialahmadi, A. Sahebkar,
Cyclodextrins: Potential therapeutics against atherosclerosis, Pharmacol. Ther. 214 (2020)
107620. https://doi.org/https://doi.org/10.1016/j.pharmthera.2020.107620.
[151] T. Irie, K. Uekama, Cyclodextrins in peptide and protein delivery, Adv. Drug Deliv. Rev.
36 (1999) 101–123. https://doi.org/10.1016/S0169-409X(98)00057-X.
[152] L. Illum, Nasal drug delivery: New developments and strategies, Drug Discov. Today. 7
56
(2002) 1184–1189. https://doi.org/10.1016/S1359-6446(02)02529-1.
[153] E.M.M. Del Valle, Cyclodextrins and their uses: A review, Process Biochem. 39 (2004)
1033–1046. https://doi.org/10.1016/S0032-9592(03)00258-9.
[154] X. Ma, Y. Zhao, Biomedical Applications of Supramolecular Systems Based on Host-
Guest Interactions, Chem. Rev. 115 (2015) 7794–7839.
https://doi.org/10.1021/cr500392w.
[155] E. Pinho, M. Grootveld, G. Soares, M. Henriques, Cyclodextrins as encapsulation agents
for plant bioactive compounds, Carbohydr. Polym. 101 (2014) 121–135.
https://doi.org/10.1016/j.carbpol.2013.08.078.
[156] Y. Zhou, G. Cheng, K. Chen, J. Lu, J. Lei, S. Pu, Adsorptive removal of bisphenol A,
chloroxylenol, and carbamazepine from water using a novel β-cyclodextrin polymer,
Ecotoxicol. Environ. Saf. 170 (2019) 278–285.
https://doi.org/10.1016/j.ecoenv.2018.11.117.
[157] B. Chen, S. Chen, H. Zhao, Y. Liu, F. Long, X. Pan, A versatile Β-cyclodextrin and
polyethyleneimine bi-functionalized magnetic nanoadsorbent for simultaneous capture of
methyl orange and Pb(II) from complex wastewater, Chemosphere. 216 (2019) 605–616.
https://doi.org/10.1016/j.chemosphere.2018.10.157.
[158] S. Abdou, J. Collomb, F. Sallas, A. Marsura, C. Finance, Beta-cyclodextrin derivatives as
carriers to enhance the antiviral activity of an antisense oligonucleotide directed toward a
coronavirus intergenic consensus sequence, Arch. Virol. 142 (1997) 1585–1602.
https://doi.org/10.1007/s007050050182.
[159] E. De Clercq, Antiviral therapy for human immunodeficiency virus infections, Clin.
Microbiol. Rev. 8 (1995) 200–239. https://doi.org/10.1128/cmr.8.2.200.
[160] T. Moriya, H. Kurita, K. Matsumoto, T. Otake, H. Mori, M. Morimoto, N. Ueba, N.
Kunita, Potent Inhibitory Effect of a Series of Modified Cyclodextrin Sulfates on the
Replication of HIV-1 in Vitro, J. Med. Chem. 35 (1992) 4767.
https://doi.org/10.1021/jm00103a016.
[161] T. Moriya, K. Saito, H. Kurita, K. Matsumoto, T. Otake, H. Mori, M. Morimoto, N. Ueba,
N. Kunita, A New Candidate for an Anti-HIV-1 Agent: Modified Cyclodextrin Sulfate
(mCDS71), J. Med. Chem. 36 (1993) 1674–1677. https://doi.org/10.1021/jm00063a018.
[162] T. Otake, D. Schols, M. Witvrouw, L. Naesens, H. Nakashima, T. Moriya, H. Kurita, K.
Matsumoto, N. Ueba, E. De Clercq, Modified cyclodextrin sulphates (mCDS11) have
potent inhibitory activity against HIV and high oral bioavailability, Antivir. Chem.
Chemother. 5 (1994) 155–161. https://doi.org/10.1177/095632029400500303.
57
[163] R. Anand, S. Nayyar, J. Pitha, C.R. Merril, Sulphated Sugar Alpha-Cyclodextrin Sulphate,
a Uniquely Potent Anti-HIV Agent, Also Exhibits Marked Synergism with AZT, and
Lymphoproliferative Activity, Antivir. Chem. Chemother. 1 (1990) 41–46.
https://doi.org/10.1177/095632029000100107.
[164] A. Leydet, C. Moullet, J.P. Roque, M. Witvrouw, C. Pannecouque, G. Andrei, R. Snoeck,
J. Neyts, D. Schols, E. De Clercq, Polyanion inhibitors of HIV and other viruses. 7.
Polyanionic compounds and polyzwitterionic compounds derived from cyclodextrins as
inhibitors of HIV transmission, J. Med. Chem. 41 (1998) 4927–4932.
https://doi.org/10.1021/jm970661f.
[165] L. Long, Y. Xue, X. Hu, Y. Zhu, Study on the influence of surface potential on the nitrate
adsorption capacity of metal modified biochar, Environ. Sci. Pollut. Res. 26 (2019) 3065–
3074. https://doi.org/10.1007/s11356-018-3815-z.
[166] M.M. Hoog Antink, S. Beutel, K. Rezwan, M. Maas, Tailoring electrostatic surface
potential and adsorption capacity of porous ceramics by silica-assisted sintering,
Materialia. 12 (2020) 100735. https://doi.org/10.1016/j.mtla.2020.100735.
[167] M.I. Cedergren, A.J. Selbing, O. Löfman, B.A.J. Källen, Chlorination byproducts and
nitrate in drinking water and risk for congenital cardiac defects, Environ. Res. 89 (2002)
124–130. https://doi.org/10.1006/enrs.2001.4362.
[168] W.A.M. Hijnen, E.F. Beerendonk, G.J. Medema, Inactivation credit of UV radiation for
viruses, bacteria and protozoan (oo)cysts in water: A review, Water Res. 40 (2006) 3–22.
https://doi.org/10.1016/j.watres.2005.10.030.
[169] V. Poornima Parvathi, M. Umadevi, R. Sasikala, R. Parimaladevi, V. Ragavendran, J.
Mayandi, G. V. Sathe, Novel silver nanoparticles/activated carbon co-doped titania
nanoparticles for enhanced antibacterial activity, Mater. Lett. 258 (2020) 126775.
https://doi.org/10.1016/j.matlet.2019.126775.
[170] T. Motomura, Y. Miyashita, T. Ohwada, M. Onishi, N. Yamamoto, et al Motomura, T.M.
Motomura Yuko; Ohwada, Takashi; Onishi; Makoto; Yamamoto, Naoki., Material for
removing HIV and its related substances, Patent US5667684A, 1997.
[171] M. Kadhom, B. Deng, Metal-organic frameworks (MOFs) in water filtration membranes
for desalination and other applications, Appl. Mater. Today. 11 (2018) 219–230.
https://doi.org/10.1016/j.apmt.2018.02.008.
58
There is no conflict of interest.
This submission has not been published previously and not under consideration for publication
elsewhere. If it will be published, it will not be published elsewhere in the same form, in English
or in any other language, including electronically without the written consent of the copyright-
holder.
Highlights
-Type of viruses, their classification and physicochemical properties are reviewed
-A detailed review of different adsorbents for virus inactivation is reported
-Potential of biological materials to adsorb virus is discussed
- Perspectives of adsorption research for virus inactivation are reported

File (1)

Content uploaded by Lotfi Sellaoui
Author content
ResearchGate has not been able to resolve any citations for this publication.
ResearchGate has not been able to resolve any references for this publication.