Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

197

Emerging Nanofibrous Air Filters for PM2.5 Removal

6.1 Particulate Matter


Atmospheric fine particulate matter (PM) possesses a large number of hazardous
and noxious oxides of sulfur, nitrogen, and carbon, heavy metal, microorgan-
isms, etc. Atmospheric PM has diverse sources, including suspended dust,
high-temperature metallurgical processes, atmospheric reactions, and various
incomplete combustion activities (e.g. vehicular emission, coal combustion,
biomass burning, diverse industrial combustion processes) [1–6]. PM is the dead-
liest form of air pollution and has severely influenced human being’s living quality.
PM is commonly classified into PM2.5 and PM10 with an aerodynamic diameter
below 2.5 and 10 μm [7, 8], respectively. Particularly, PM2.5 tends to penetrate
deeply into gas exchange regions of the lung and blood, is deadly, and induces
serious human health concerns, including heart disease, lung infections, and even
cancer [7, 8].
In the late 1980s and early 1990s, PM2.5 was first publicly discussed, mainly
involving its composition and health impact [9, 10]. In the latest decade, PM2.5
has become a notorious air pollutant across the globe. Many Asian and some
European and American cities have been experiencing severe PM pollution asso-
ciated with rapid industrialization, energy consumption, urbanization, and popu-
lation growth [11–16]. For example, the Indian subcontinent, as one of the rapidly
developing middle-income Asian countries, has recorded 1.09 million deaths per
year due to PM2.5 , according to Cohen’s study in 2017 [17]. The latest global
burden of disease (GBD) estimates that India’s population-weighted mean PM2.5
concentration reached 74.3 μg m−3 , much higher than that required by the World
Health Organization (WHO) (PM2.5 < 10 μg m−3 ) [17].
PM2.5 can stay stable in the air between hours and weeks owing to their small
sizes. Stable PM2.5 with high concentration results in a hazy day (Figure 6.1). They
can scatter visible light and reduce visibility due to the similarity between their
particle size and visible light wavelengths [19]. PM2.5 particles in heavily polluted
air can have very complicated compositions, typically having carbon-derived
matters (e.g. carbon dioxide and carbon monoxide), silicates (SiO3 2− ), sulfur
dioxide (SO2 2− ), sulfates (SO4 2− ), silicon dioxide (SiO2 ), nitrates (NO3 − ),
ammonium (NH3 ), chloride, and biological matters (e.g. bacteria and viruses)
[1–3, 18, 20–22]. Meanwhile, owing to the existence of various ions and water

Artificially Intelligent Nanomaterials for Environmental Engineering, First Edition.


Peng Wang, Jian Chang, and Lianbin Zhang.
© 2019 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2019 by Wiley-VCH Verlag GmbH & Co. KGaA.
198 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

(a) (b)

Figure 6.1 (a) Photograph of Beijing on a sunny day. (b) Photograph of Beijing on a hazy day
with hazardous PM2.5 level. Source: Liu et al. 2015 [18]. Reprinted with permission of Springer
Nature.

vapor in the atmospheric environment, it is likely that high polar functional


group is largely distributed onto the surface of the PM particles, leading to their
high polarities in the air [18, 23].

6.2 Traditional Technology


During hazy days, ventilation system and central air conditioning are com-
monly used to filter out PM and produce fresh air. However, this technology
is only available in modern commercial buildings [24, 25], and it entails high
energy consumption to power bulk pumping systems. Therefore, air filter with
natural ventilation is regarded as a more ideal method to obtain fresh air because
no additional energy input required and it suits both outdoor and indoor
application scenarios [18].
Recently, existing air filter technology has gradually become a popular strategy
to remove PM2.5 from polluted air. There are two typical examples of air filters
in common use, namely, porous membrane filter and fibrous air filter, respec-
tively [18].
Similar to a water filtration membrane, porous membranes fabricated by cre-
ating pores on a solid substrate are conventionally applied for atmospheric PM
particle removal (Figure 6.2a). However, conventional porous membrane usually
has extremely small pore size to filter out PM2.5 , and the porosity of this type of
air filter is low (<30%). Hence, it would result in an extremely low airflow across
these membrane air filters.
As a strategy to reduce the air pressure drop, fibrous membranes, consisting of
stacked fibers of various forms, have been widely applied in recent commercial air
filters, such as disposable respirators, industrial gas cleaning equipment, clean-
room air purification systems, automotive cabin air filters, and indoor air puri-
fiers [26]. The fibrous membranes are conventionally made of various materials,
such as melt-blown fibers, glass fibers, and spun-bonded fibers [27]. Compared
with the porous membranes, the conventional microfibrous membranes provide
6.3 Nanofibrous Membrane Air Filters 199

Airflow Airflow

Small opening
size exclusion
Thick
Micron size
Porous membrane air filter Commercial fibrous air filter
(a) (b)

Figure 6.2 (a) Schematic illustration of porous air filter capturing PM particles by size
exclusion. (b) Schematic illustration of bulky fibrous air filter capturing PM particles by thick
physical barrier and adhesion. Source: Liu et al. 2015 [18]. Reprinted with permission of
Springer Nature.

higher air permeability with more than 70% porosity, but their PM2.5 removal
performance is still not satisfactory due to the microsized diameter of the indi-
vidual fibers, ranging from several microns to tens of microns [18, 26, 28–30].
Therefore, large fiber diameter, uncontrollable porosity, and packing density of
fibrous air filter cause the several deficiencies, including the bulkiness, material
consumption, nontransparency, and the trade-off between airflow and filter effi-
ciency (Figure 6.2b) [18, 27].

6.3 Nanofibrous Membrane Air Filters


Nanofibrous membrane possesses fiber diameters in the range of 10–1000 nm,
leading to a multitude of attractive features, such as high specific surface area,
extremely large aspect ratio, high porosity, low resistance to airflow, and easy
functionalization ability. All of these features are beneficial for PM2.5 removal
(Figure 6.3) [32–37].

6.3.1 Filtration Mechanism


The filtration mechanism determines the PM retention efficiency on air filter. Dif-
fering from conventional filtration media, the nanofibrous air filter can physically
capture tiny particles much smaller than the pore size of the filter, which is beyond
simple physical sieving or interception [38].
With the understanding of high polarity of PM in the air as mentioned above,
nanofibrous membranes with high polarity have been made to aim at high
adhesion interactions between PM particles and the nanofibers, including elec-
trostatic interaction, dipole–dipole and dipole-induced intermolecular forces,
coordination interaction, and hydrogen bonding [18, 23]. Based on that, it could
be inferred that polymeric air filters with higher dipole moment of their repeating
unit, such as polyimide (PI) (6.16 D) [39], Nylon-6 (3.67 D) [40], and polyacry-
lonitrile (PAN) (3.6 D), could possess stronger dipole–dipole and dipole-induced
200 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

High airflow

High transparency

High adhesion force

Figure 6.3 Schematic illustration of nanofibrous membrane air filter with high transparency
and low resistance to airflow that captures PM particles by strong surface adhesion. Source:
Chang et al. 2018 [31]. Reprinted with permission of Royal Society of Chemistry.

intermolecular forces with PM, resulting in high PM capture efficiencies. Mean-


while, some functional groups from polymers such as —OH, —COOH, —NH3 ,
and —C=O can induce hydrogen-bonding and/or electrostatic interactions.
As known, metal–organic framework (MOF), as porous crystalline materials,
consists of metal ions or clusters and organic links [41–43] and possesses such
advantages as large surface area, multifunctions, and high thermal stability.
Recently, MOF has been employed to bond PM on the active metal sites via
coordination bonding. The unbalanced metal ions on the surface of MOFs offer
positive charges, which can polarize the surface of PM and thus enhance the
electrostatic interaction between MOF nanocrystal and PM [23].
The PM capture using nanofibrous air filter at different stages is shown in
Figure 6.4a–d. As PM particle capturing continues, the incoming particles would
directly attach onto and merge together with the pre-existing PM along the
fibers, leading to stable aggregates around the nanofibers [18]. As shown in
Figure 6.4e,f, The PM particle being captured by nanofibers would wrap around
the fibers tightly, deform, and finally reach to a stable spherical shape. This
phenomenon is to the benefit of PM removal since it allows PM particles to
enlarge their contact areas with nanofibers and thus to bind tightly onto the
nanofibers, ensuring a good capture performance.
Moreover, humidity has also been taken into consideration. Humidity could
help nanofibers to achieve better PM capture since the ambient water content
increased the capillary force between the PM particles and nanofibers during PM
attachment [18].

6.3.2 Fabrication Methods


Guided by the above interaction principles, nanofibrous air filters have been
made for PM2.5 removal from suitable polymers, including PAN [18, 44–46],
polyvinylpyrrolidone (PVP) [18, 47], PS [18], polyvinyl alcohol (PVA) [18, 48], PI
[39], PU [49, 50], poly(lactic acid) (PLA) [51], poly(m-phenylene isophthalamide)
6.4 Applications 201

Merge to
existing PM

her
get
Attach to rge to
nanofiber me
PM
(a) (b) (e)

(f)
(c) (d)

Figure 6.4 (a–d) Schematics showing the mechanism of PM capture by the nanofibrous filter
at different time sequences. (e) SEM image showing the morphologies of attached PM that
formed a coating layer wrapping around the nanofibers (scale bar: 1 μm). (f ) SEM image
showing that the nanofiber junction has more PM aggregated to form bigger particles (scale
bar: 1 μm). Source: Liu et al. 2015 [18]. Reprinted with permission of Springer Nature.

[52], polyethylene oxide (PEO) [53], PC [54, 55], silk [56], Nylon-6 [40], Nylon-66
[57], PMMA [47], and protein [58, 59], as well as composite materials, involving
PVDF doped with negative ions powder [36], PVDF/polytetrafluoroethylene
(PTFE) [60], PVC/PU [61], PAN/fluorinated PU [62], PAN/silica [63], Nylon-6/
PAN [27], Nylon-6/boehmite [64], PVA/PAN [65], PAN/ionic liquid [66], polysul-
fone/titania (TiO2 ) [67], PAN/polysulfone [68], PVA/TiO2 [69], PVDF/Ag/Al2 O3
[70], PMMA/cyclodextrin [71], PS/cyclodextrin [72], PC/benzyl triethylammo-
nium chloride [73], PLA/TiO2 [74], PAN/MOF [23], etc.
Up to now, a range of techniques such as template synthesis, electrospin-
ning, chemical vapor deposition (CVD), drawing, self-assembly, solvothermal
synthesis, phase separation, etc. have been explored to fabricate nanofibrous
membranes [18, 75–81].
Among these techniques, electrospinning is regarded as a simple and versatile
approach for fabrication of uniform nanofibers with diameters from micrometers
to nanometers using diverse polymer solutions [34, 82–85]. In the electrospin-
ning process, strong electrostatic force produced by a high voltage overcomes
the surface tension of the polymer solution or melt, and its fluid jet deformed
into a conical shape (Taylor cone) is ejected to the counter electrode, becoming
polymer fibers after evaporating [86–90]. Electrospun nanofibrous membranes
possess attractive properties that make them desirable in air filter area [86, 88, 89].
Recently, various new technologies have been introduced for continuous fabri-
cation of large-area filters, such as roll-to-roll blow-spinning technique [47] and
roll-to-roll hot-pressing methods [91].

6.4 Applications
This rest of this chapter reviews major milestones in this budding field of
nanofibrous air filters for PM2.5 removal. We highlight the significant recent
advances of nanofibrous membrane air filters with integrated multifunctions,
202 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

such as transparency, high thermal stability, intelligent thermal management,


self-powering, antibacterial property, toxic gases and oil removal, etc.

6.4.1 Transparent Air Filter


In 2015, Cui and coworkers pioneered the design of nanofibrous air filter by uti-
lizing electrospun polymeric nanofibrous membranes including PAN, PVP, PS,
PVA, and PP for indoor air protection [18]. As shown in Figure 6.5a, these elec-
trospun polymer nanofibers can lie across the holes of the metal-coated window
screen mesh without any additional energy input to push the airflow through

Window screen with metal coating

N
N O OH
n n n n n
V (3.6 D) (2.3 D) (0.7 D) (1.2 D) (0.6 D)
(a) (b) PAN PVP PS PVA PP

(c)

(d)
PM2.5 PM10–2.5
1.0
Removal efficiency

0.8
0.6
0.4
0.2
0.0
(f)
N
P
PS

C VA
C on

r
PP
pe
PV
PA

b
P

op
ar

(e)

Figure 6.5 (a) Schematic illustration of the fabrication of transparent air filter via
electrospinning methods. (b) Molecular model and formula of different polymers including
PAN, PVP, PS, PVA, and PP with calculated dipole moments of the repeating units of each
polymer. (c) SEM images of PAN, PVP, PS, PVA, and PP transparent filters before filtration.
(d) SEM images of PAN, PVP, PS, PVA, and PP transparent filters after filtration showing the PM
attachment. Scale bars in c,d: 5 μm. (e) Removal efficiency comparison between PAN, PVP, PS,
PVA, PP carbon and copper transparent filters with same fiber diameter of ∼200 nm and same
transmittance of ∼70%. Error bar represents the standard deviation of three replicate
measurements. (f ) Demonstration of using transparent filter to shut off PM from the outdoor
(right bottle) from entering the indoor (left bottle) environment. Source: Liu et al. 2015 [18].
Reprinted with permission of Springer Nature.
6.4 Applications 203

the membrane filter. The molecular models and formulae of these polymers are
shown in Figure 6.5b, corresponding to their different polarity with the dipole
moments of 3.6, 2.3, 0.7, 1.2, and 0.6 D for the repeating units of PAN, PVP, PS,
PVA, and PP, respectively. Therefore, these fibrous membranes are able to cap-
ture PM, as shown in Figure 6.5c,d. Among them, the PAN nanofibers showed the
greatest efficiency of PM2.5 removal (>95%) (Figure 6.5e) and 90% transparency
under hazardous air quality conditions (PM2.5 mass concentration > 250 μg m−3 ).
The PM2.5 capture was ascribed to the dipole–dipole and dipole-induced inter-
actions in this work, suggesting that polymers with higher dipole moment would
have better removal efficiencies of PM particles. As shown in Figure 6.5f, the left
bottle was still clear after filtering out PM, which demonstrated the PM removal
performance of the PAN transparent filter.

6.4.2 Air Filter for High Thermal Stability


Additionally, the direct removal of PM at a higher temperature, which are
produced from high-temperature processes like burning, industrial exhaust,
biomass, vehicle exhaust, and coal combustion [92–96], is another challenge
for the abovementioned polymers. In 2016, Cui’s group further developed
transparent and high-temperature stable PI nanofibrous air filter, which showed
great mechanical properties, thermal stability, and high and consistent PM2.5
removal performance (>99.5%) at the temperatures ranging from 25 to 370 ∘ C
for more than 120 hours [39]. As shown in Figure 6.6, PI filter can effectively
remove all kinds of particles (PM sizes, 0.3–10 μm) from the car exhaust gas
with high temperature ranging from 50 to 80 ∘ C. The PM concentrations after
filtration were decreased to almost the same as that of clean ambient air.
Following this concept, Zhao and coworkers utilized PI nanofibers as building
blocks to construct hierarchically porous aerogels through freeze drying and
thermally induced cross-linking (Figure 6.7) [97]. The as-prepared polyimide
nanofiber aerogels possess outstanding flexibility and toughness, ultralow den-
sity (4.6–13.1 mg cm−3 ), high porosity (99.0–99.6%), low thermal conductivity

100
Removal efficiency (%)

80

60

40

20

0
0.3 0.5 1.0 2.5 5.0 10
(a) PM size (μm) (b)

Figure 6.6 (a) Removal efficiency of PM particles from car exhaust gas. (b) PM number
concentration measurement of car exhaust with air filter. The inset shows a stainless steel pipe
coated with a PI filter. Source: Zhang et al. 2016 [39]. Reprinted with permission of American
Chemical Society.
204 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

PMDA
1) Electrospinning Homogenizing

2) Thermal
ODA imidization
Poly(amic acid)
Polyimide nanofiber
Nanofiber dispersion
membrane

Freezing

Thermal Freeze
treatment drying
Chemical Physical
cross-linking entanglement

PINFAs Uncross-linked PINFAs Frozen dispersion

Figure 6.7 Schematic illustration of the fabrication process of PI nanofiber membranes


(PINFMs). Yellow fibers represent PI nanofibers. A plausible chemical structure after
intermolecular condensation in PINFAs is presented. Source: Qian et al. 2018 [97]. Reprinted
with permission of Royal Society of Chemistry.

(29.7–31.8 mW m−1 K−1 ), and high-temperature stability (300 ∘ C for 240 hours
in air) in different dimensions. With the combination of the structural features
of the aerogels and the specific physicochemical property of PI, the filtration
efficiency remained above 99.9% under the mass concentration of PM2.5 above
200 μg m−3 for 22 hours at a velocity of 0.25 m s−1 . The calculated sorption
dynamics was about 4.3 μg g−1 s−1 .
In addition, to mitigate fire and explosion risk in operation with combustible
filtrates, in 2018, Cui et al. further developed a multifunctional nanofibrous air
filter that not only efficiently captured PM but also possessed a flame retardant
design [98]. This multifunctionality design was achieved through fabricating an
electrospun core–shell nanofibrous membrane consisting of the polar polymer
Nylon-6 as the shell and the flame retardant triphenyl phosphate (TPP) as the
core (Figure 6.8a). The Nylon-6 polymeric shell ensured the high capture effi-
ciency of 99.00% for PM2.5 owing to its large dipole moment initiating strong
binding between the fiber surface and polar PM. In terms of the flame retardant
core, TPP, it could be released through the melted Nylon-6 shell during combus-
tion of the flammable filtrate and simultaneously suppress the fire by scavenging

H and • OH free radicals generated from combustion (Figure 6.8b–e) [99]. As
a result, the self-extinguishing time of the filtrate-attached air filter was nearly
instantaneous of 0 s g−1 compared with 150 s g−1 for Nylon-6 alone.

6.4.3 Air Filter for Thermal Management


In 2017, for the first time, thermal management was introduced into nanofibrous
air filter face mask by Cui and coworkers for personal cooling/warming purposes
[100]. In this design, the nanoPE was chosen as a supporting substrate due
to its transparency to the mid-infrared (IR) radiation, and electrospun nylon
nanofibers were modified on the PE substrate. As-prepared PE/nylon composite
6.4 Applications 205

Shell: O
Syringe Nylon-6: P
O O Heat Phosphorus-containing
TPP: O
TPP free radicals (·PO)

PO H HPO
O
Core: TPP Nylon-6: PO OH HPO2
N n
H
Collec
(a) tor (b) (c)

Ignition Flame
Ignition Flame and sources
sources Airflow retarded
Airflow explosions

TPP
Nylon-6 @Nylon-6
(d) fiber (e) fiber

Figure 6.8 (a) Schematic illustration for the fabrication of the nanofibers by electrospinning.
(b) Molecular structure of TPP and Nylon-6. (c) The TPP exhibits flame retardancy by
scavenging • H and • OH free radicals during combustion. (d) The air filtration material
composed of Nylon-6 nanofibers. Combustive dusts accumulated on the filter are easily
ignited by ignition sources, leading to dust explosions. (e) The working mechanism of the
multifunctional air filter. Source: Liu et al. 2018 [98]. Reprinted with permission of American
Chemical Society.

membrane mask exhibited high PM capture efficiency (99.6% for PM2.5 ) with low
pressure drop. As shown in Figure 6.9, it showed excellent heat dissipation and
high IR transparency (92.1%), generating the radiative cooling effect. Separately,
if the nanoPE substrate was coated by a thin layer of Ag before the nylon fiber
deposition, it gave rise to a high IR reflectance (87.0%) for personal warming
purpose. These two type of face masks are desirable for personal thermal comfort
under hot and cold weather, respectively.

Cooling:

Nylon-6 fi
ber
Fiber/
nanoPE

E
noP Warming:
che d na
Pun
°C
Nanopore
37.5
Nanofiber Fiber/Ag
/nanoPE
23.5
(a) (b)

Figure 6.9 Face mask consisting of Nylon-6 nanofibers on top of needle-punched


nanoporous-polyethylene (nanoPE) substrate (a) and thermal imaging of faces covered with
the sample (b). Source: Yang et al. 2017 [100]. Reprinted with permission of American
Chemical Society.
206 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

6.4.4 Air Filter for Mass Production


For the purpose of fast and large-scale commercial production of air filter, Cui
and coworkers proposed and tested a roll-to-roll transfer fabrication method of
nanofibrous air filter (Figure 6.10a) [40]. Accordingly, an ameliorated electrospin-
ning design was built based on fast transfer of electrospun Nylon-6 nanofiber
film from rough copper foil to a rolling plastic mesh substrate by laminating and
peeling method. Compared with the conventional electrospinning method, the
transfer method is 10 times faster and has better filtration performance at the
same transmittance (>99.97% removal of PM2.5 at ∼73% of transmittance).
Later, a roll-to-roll blow-spinning technique was developed by Cui and Wu
et al. and applied to the mass production of transparent air filters (Figure 6.10b)
[47]. By sequential multi-needle blow spinning, such transparent nanofibrous
air filter films could be coated rapidly on regular window screens (Figure 6.10c),
which acquired 90.6% PM2.5 removal efficiency over 12 hours under hazy air
conditions (PM2.5 mass concentration > 708 μg m−3 ). Moreover, PM can be easily
removed by gentle wiping (Figure 6.10d).

6.4.5 Self-Powered Air Filter


To endow the nanofibrous air filters with more intelligence, Wang et al. demon-
strated a concept of self-powered air filter for capturing PM from automobile
exhaust using triboelectrification effect. The triboelectric nanogenerator (TENG)
in this work harvested electric energy from mechanical movements. In 2015,
they fabricated a TENG based on the collision between PTFE pellets and elec-
trode plates that formed a space electric field as high as 12 MV m−1 , and thus

Airflow

(b)

(a) (d) (c)

Figure 6.10 (a) Photograph of a roll-to-roll process for the transferring of electrospun
nanofiber film onto a plastic mesh in a continuous fabrication process for PM2.5 filter. Source:
Xu et al. 2016 [40]. Reprinted with permission of American Chemical Society. (b) Schematic
illustration of the blow-spinning method of the window screen coating for indoor protection.
(c) Real window consisting of a removable metallic screen coated with PAN blow-spun fibers.
(d) Successful wiping of nanofibers from the window screen using a tissue paper. Source:
Khalid et al. 2017 [47]. Reprinted with permission of American Chemical Society.
6.4 Applications 207

PM could be absorbed under the electrostatic force, generating 95.5% removal of


PM2.5 [101].
In 2017, they developed a TENG-assisted positively charged PI electrospun
nanofibrous air filter to enhance the removal of especially superfine PM with
a diameter smaller than 100 nm [102]. As shown in Figure 6.11a,b, an electric
field forms around the stainless steel meshes and PI nanofiber film, and thus
triboelectrification-induced electrostatic absorption effect can work on a larger
particle size span, from nanoparticles to microparticles. As a result, the great-
est enhancement of PM removal efficiency was 207.8% at the particle diameter of
76.4 nm, and the highest removal efficiency was 90.6% at the diameter of 33.4 nm.
This technology with zero ozone release and low pressure drop offers a great
promise in air cleaning and haze treatment.
In the same year, Ko and coworkers fabricated a percolation network of Ag
nanowire on nylon mesh as a transparent, reusable, and active PM2.5 air filter
[103]. As shown in Figure 6.11c, by applying a low voltage on the Ag nanowire
network, the membrane exhibited a high PM2.5 removal efficiency (>99.99%)
due to voltage-induced strong electrostatic force. Meanwhile, the obtained air
filter was transparent, low power consumption, antibacterial, and reusable after
simple washing.
In 2018, Wang and coworkers prepared a washable high-efficiency triboelec-
tric air filter (TAF) for efficiently removing the PMs [104]. The TAF consists

Electrostatic zone

Dust air Dust air Clean air

Pl nanofibers
Pl nanofibers +R-TENG
(a) (b)
Transparent,
reusable, active
PM filter

Ag NW
percolation
PM network
Polluted air Filtrated air

(c) + Voltage

Figure 6.11 Schematic illustration of the filtration mechanism. (a) Without R-TENG. (b) With
R-TENG. Source: Gu et al. 2017 [102]. Reprinted with permission of American Chemical Society.
(c) Schematic illustration of the filtration mechanism of the Ag nanowire-coated filter. Source:
Jeong et al. 2017 [103]. Reprinted with permission of American Chemical Society.
208 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

Severely polluted Good

μg m−3 μg m−3

Heavily polluted Excellent

μg m−3 μg m−3

Figure 6.12 The schematic diagram of the measurement of the removal efficiency of the face
mask made of the TAF. The face mask was worn by a man for four hours. The concentration of
PM2.5 changed from severely polluted to good and from heavily polluted to excellent. Source:
Bai et al. 2018 [104]. Reprinted with permission of John Wiley and Sons.

of multilayers of the PTFE and nylon fabrics, which can be charged by simply
rubbing the PTFE and nylon fabrics against each other. The removal efficiency
of the charged TAF can be greatly enhanced by introducing the electrostatic
attraction in removing the PMs. After charging, the removal efficiencies of PM0.5
and PM2.5 were increased from 26.3% to 84.7% (3.22 times increase) and 69.1%
to 96.0% (1.39 times increase), respectively. As shown in Figure 6.12, the PM2.5
filtering through the worn face mask decreased from 266.71 μg m−3 (severely
polluted condition) to 54.71 μg m−3 (good air condition), and correspondingly,
PM2.5 in heavily polluted condition with the concentration of 192.32 μg m−3
could be decreased to 27.93 μg m−3 , representing an excellent air condition. In
addition, the TAF can be easily cleaned with commercial detergent, and the
removal efficiency was maintained after five washing cycles.

6.4.6 Nanofibrous Air Filter for the Simultaneous Removal of PM


and Toxic Gases
In reality, the pollution of PM particles is always concomitant with gaseous
chemicals, such as formaldehyde (HCHO), sulfur dioxide (SO2 ), nitrogen dioxide
(NO2 ), benzene, dioxin, ozone, etc. [3, 22, 59, 105, 106]. Taking consideration
of these issues, MOFs have offered some help in the capture of harmful gas
[107–113] and adsorption and degradation of chemical toxic agents [114, 115].
In 2016, for the first time, Wang and coworkers prepared MOF-based nanofi-
brous air filter for the simultaneous removal of PM and toxic gases [23]. In this
work, MOFs (ZIF-8, UiO-66-NH2 , MOF-199, Mg-MOF-74) were embedded
within polymers (PAN, PS, PVP) to prepare nanofibrous air filters by the
electrospinning method, and among all, UiO-66-NH2 /PAN and MOF-199/PAN
hybrid nanofibrous air filters showed the best PM removal performance and SO2
adsorption (Figure 6.13). This is because the polar functional groups and positive
6.4 Applications 209

92 24
0.020 SO2 dynamic adsorption capacity

SO2 dynamic adsorption


PM2.5
Removal efficiency (%)

Pressure drop

Pressure drop (Pa)


88 PM10
0.015 18

capacity (g/g)
84
0.010 12

80
0.005 6

76
0.000 0
N N N N N N N N N N N
PA /PA 9/PA /PA 4/PA 8/PA PA 8/PA 4/PA 9/PA /PA
O3 9 H 2 -7 - - 7 9 H2
Al 2 OF-1 -66-N MOF ZIF ZIF OF- OF-1 66-N
M iO - -M M -
(a) U Mg (b) Mg UiO

Figure 6.13 (a) PM removal efficiency of PAN filter, Al2 O3 /PAN filter, and PAN/MOF filters
tested on hazy days in Beijing (T = 23.4 ∘ C, RH = 58.6%, PM2.5 = 350 μg m−3 ,
PM10 = 720 μg m−3 ). (b) The dynamic adsorption capacities of SO2 on PAN filter and PAN filters
with different MOF materials at 25 ∘ C with a 100 ppm of SO2 /N2 flow at the rate of 50 ml min−1 .
Source: Zhang et al. 2016 [23]. Reprinted with permission of American Chemical Society.

charges on the surface of MOFs can generate electrostatic and polarity-induced


interaction with PMs to bond them tightly, while functionalities such as amines
and open metal sites of hybrid fibers are crucial for kinetic adsorption of acidic
polar gas species.
Furthermore, Wang and coworkers used a roll-to-roll hot-pressing method
and fabricated MOF-based air filters [91]. In this method, the MOF nanocrystals
were generated and stably immobilized onto the flexible substrates via continu-
ously roll-to-roll pressing between two rollers with high temperature for several
times (Figure 6.14). The MOF-based nanofibrous air filters showed long-term

Catalyst
pulverization

MOF
Substrate MOF Particulate matter
Power plant

MOF@Plastic mesh
Refinery (80 ~ 100 °C)

MOFilters
Traffic MOF@Melamine foam MOF@Metal mesh
(~150 °C) (>300 °C)

Biomass
combustion
MOF@Nonwoven fabric MOF@Glass cloth
Roll-to-roll production (150–250 °C) (>300 °C)

Figure 6.14 The schematic representation of the roll-to-roll production of various MOF-based
filters for PM removal. Source: Chen et al. 2017 [91]. Reprinted with permission of John Wiley
and Sons.
210 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

(i.e. 30 consecutive days) and consistently high PM removal (>90%) under a


wide temperature range (80–300 ∘ C). Moreover, such air filter can be easily
cleaned by washing with water and ethanol and reused for three times without
any change of structure, morphology, and PM removal efficiency. This air filter
can be applied in the waste gas treatment in existing piping systems, vehicle or
aircraft engine pipes, and reaction vessels.
Further, Kim and coworkers proposed heterogeneous nucleation-assisted hier-
archical growth of flowerlike MOFs (Zn-based zeolite imidazole frameworks,
ZIF-L) (Figure 6.15a) on several substrates (e.g. glass, PU foam, nylon microfibers,
and PP microfibers), as air filters for efficient PM removal [116]. MOFs assembled
on PP microfibers showed superior properties in terms of PM2.5 removal effi-
ciency (92.5 ± 0.8% for PM2.5 and 99.5 ± 0.2% for PM10 ) (Figure 6.15b), pressure
drop (10.5 Pa at 25 l min−1 ), long-term stability, and recyclability even after wash-
ing. The PM filtering performances are mainly ascribed to the unique 2D-shaped
structure anchored on individual fibers and electrostatic PM screening imparted
by the partially charged surface of MOFs.
In 2018, Zhang and coworkers presented an immersion method for embedding
the MOFs into electrospun nanofibers, which served to effectively remove both
PM2.5 and formaldehyde (Figure 6.16a) [117]. The PM2.5 filtration efficiency was
raised from 74.5% to 87.2%, after integrating ZIF-67 nanocrystals into the electro-
spun PAN nanofibers. Moreover, the PM2.5 filtration efficiency remained at more
than 99% and dropped by only 0.25% during one-month continuous PM2.5 filtra-
tion test (Figure 6.16b). The air filter obtained a formaldehyde removal efficiency
of 84%.
In addition to MOFs, proteins have great promise for environmentally friendly
air filtration with capability of capturing PM and toxic gases like CO and HCHO.
For instance, in 2016, Zhong and coworkers employed soy [58] and gelatin
protein [59] and fabricated electrospun nanofibrous air filter for the PM and
toxic gas removal. The main mechanism is demonstrated in Figure 6.17a. The
amine groups from gelatin and soy protein can interact with carbon oxide

PM2.5 PM10
100
PM removal (%)

90

80

70

60
0
2 μm PP ZIF-8_PP H-ZIF-L_PP

(a) (b)

Figure 6.15 (a) SEM images of hierarchical flowerlike MOFs. (b) PM removal efficiencies of PP
microfibers, ZIF-8_PP microfibers, and H-ZIF-L_PP microfibers. Source: Koo et al. 2018 [116].
Reprinted with permission of American Chemical Society.
6.4 Applications 211

ZIF-67@PAN 100 PM2.5

Filtration efficiency (%)


99

Diffusion 98
Co(AC)2/PAN
97

96

95
3 6 9 12 15 18 21 24 27 30
Date (d)
(a) (b)

Figure 6.16 (a) Schematic representation of the fabrication procedures for MOF-based PAN
filters. (b) Long-term PM2.5 filtration test results for MOF-based PAN filters. Inset image in (b) is
the photograph of a large-scale MOF-based PAN filters (length: 36 cm). Source: Bian et al. 2018
[117]. Reprinted with permission of Royal Society of Chemistry.

rface
s on su
ction al group
Rich fun

COOH COOH

R R (hydrophobic)
R: Hydrophobic OH OH
NH2 N Hydrogen bonding
Charge–charge interaction
Polar O C Hydrophobic interaction ...
H H
PM Charged/Ionic C Polar
H H Aldimine linkage
Hydrophobic
PM Charged/Ionic
(a) Formaldehyde
Hydrophobic

100% O
100%
C Formaldehyde C O Carbon monoxide
Removal efficiency (%)

Removal efficiency (%)

80% H H
80%

50% 50%
2.25 g m–2

2.80 g m–2

3.43 g m–2

3.80 g m–2

2.25 g m–2

2.80 g m–2

3.43 g m–2

3.80 g m–2
164 g m–2

164 g m–2

20% 20%
5% 5%
(b) Gelatin fibers HEPA (c) Gelatin fibers HEPA

Figure 6.17 (a) Schematic illustration of the interaction-based filtration mechanism for
soy-protein-based nanofabrics. Source: Souzandeh et al. 2016 [58]. Reprinted with permission
of American Chemical Society. (b) Formaldehyde (HCHO) removal efficiency comparison
between gelatin-based filters with different areal density and that of the commercial filter.
(c) Carbon monoxide (CO) removal efficiency comparison between gelatin-based filters with
different areal density and that of commercial filter. Source: Souzandeh et al. 2016 [59].
Reprinted with permission of Royal Society of Chemistry.
212 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

(CO) and aldehyde group in formaldehyde (HCHO) to form aldimine bonds


with the purpose of the toxic gas removal, and simultaneously PM and other
toxic chemicals can be captured via hydrogen bonding and charge–charge,
and polar–polar interactions with protein nanofibers. Consequently, the soy
protein/PVA electrospun nanofibers possessed HCHO removal efficiency
ranging from 30.0% to 62.5% and CO removal efficiency between 76.9% and
90.9%. For gelation fibers, the HCHO and CO removal efficiency could reach
as high as 80% and 76%, respectively, in comparison with less than 5% and 3%
HCHO and CO removal efficiency on commercial air filters (Figure 6.17b,c).
Furthermore, in 2018, Ke and coworkers constructed multifunctional fibrous
filter for passive room air purification [118]. As-synthesized nanocrystalline
MnO2 catalysts showed excellent HCHO catalytic activity at low temperatures
[119, 120], and further was chosen to fabricate an MnO2 /PE/PP composite
filter with PM2.5 filtration, HCHO adsorption, and catalytic abilities in air
pollutant abatement. As a consequence, as-prepared MnO2 /PE/PP composite
filter exhibited complete oxidation for HCHO within 60 minutes and acceptable
reversibility over five cycles. After corona charging, MnO2 /PE/PP composite fil-
ter kept reasonable filtration efficiency and pressure drop while at the same time
possessed fast catalytic degradation of 150 ppm HCHO within 15 minutes.

6.4.7 Nanofibrous Air Filter with Antibacterial Functions


Owing to the presence of bacteria and viruses in air pollutant, research
efforts have been also made to provide nanofibrous air filters with antibacterial
functions by using Ag [56], ZnO [77], and TiO2 [74] nanoparticles. Ag nanopar-
ticles, as well-known broad-spectrum antibacterial agents, can induce the
degeneration of protein [121, 122]. ZnO and TiO2 nanoparticles could generate
reactive hydroxyl radicals under UV irradiation, and the hydroxyl radicals, once
generated, are highly photooxidative and effectively inhibit bacterial growth by
penetrating the cell wall of the bacteria [123–127]. Therefore, these inorganic
nanoparticles have been employed to inhibit bacterial growth along with PM
removal.
In 2015, Wang and coworkers used atomic layer deposition to uniformly
seed ZnO on the surface of expanded PTFE matrix and then synthesized well-
aligned ZnO nanorods onto the PTFE substrate under hydrothermal conditions
(Figure 6.18) [77]. ZnO nanorod modified nanofibrous air filter possessed high
efficiency of PM removal (>99.9999%) and high antibacterial activity (>99.0%
sterilization rates against both Gram-positive and Gram-negative bacteria).
In 2016, Zhao and coworkers fabricated hybrid PLA/TiO2 electrospun nanofi-
brous air filter by the methods of CVD and hydrothermal synthesis, which can
effectively inhibit the propagation of bacteria [74]. In this system, the intro-
duction of TiO2 nanoparticles endows the obtained air filter with antibacterial
properties. The reactive hydroxyl radicals generated by the photocatalysis of TiO2
nanoparticles with light irradiation cause the peroxidation of the polyunsatu-
rated phospholipid of the bacteria cell membrane, leading to a loss of respiratory
activity and thus killing the bacteria [127, 128]. As a result, the obtained hybrid
PLA/TiO2 electrospun nanofibrous membrane loading with 1.75 wt% TiO2
6.4 Applications 213

Dust

PTFE PTFE

Airflow

Clean air
Polluted air

ZnO nanorods

Functionalized air filter with ZnO nanorods

Figure 6.18 Schematic of the filtration process of the ZnO-functionalized PTFE filters and the
SEM images of the filters modified with ZnO nanorods. Source: Zhong et al. 2015 [77].
Reprinted with permission of American Chemical Society.

nanoparticles exhibited a high PM removal efficiency (99.996%) with a relatively


low pressure drop (128.7 Pa), as well as a high antibacterial activity of 99.5%.
In 2015, Singh and coworkers fixed Ag into air filter to explore its antibacterial
and detoxification ability [70]. Hybrid electrospun PVDF-Ag-Al2 O3 nanofibrous
air filter was highly efficient to filter 0.36 μm particles, leading to 99.17% filtra-
tion efficiency. Silver incorporation enabled the air filter to kill the pathogens
and detoxify the chemical compounds that come into contact with them, which
provided suitable disinfection with more than 99.5% antibacterial efficiency.
In 2016, Zhang and coworkers reported a silk nanofiber air filter, which showed
a filtration efficiency of 98.8% for PM2.5 and 96.2% for 300 nm particles with a
low pressure drop [56]. In addition, Ag nanoparticles could be easily incorpo-
rated into silk nanofibers, enabling antibacterial activity into the air filter, against
Escherichia coli, a typical Gram-negative bacterium, and Staphylococcus aureus,
a typical Gram-positive bacterium [129–131].
Recently, Xiong and coworkers introduced silver nanoparticles into nanofibers
to prepare soy protein isolate (SPI)/polymide-6 (PA6)/Ag nanofibrous air filter,
which exhibited over 95% filtration efficiency of PM (PM size less than 0.3 μm)
and prevented growth of microorganisms (E. coli and Bacillus) over the filter
media [132].
In addition to these metal nanoparticles, carbon nanotubes (CNTs) are also
antibacterial candidates to cause damages of bacterial cells, which penetrate into
the interior of the bacteria cell and affect the cell division process [133–135]. In
2015, Zhong and coworkers created multiwalled carbon nanotubes (MWCNTs)
on a porous alumina ceramic membrane via CVD method [75]. The hybrid filters
exhibited high PM filtration efficiency and antibacterial property. As shown in
Figure 6.19, the presence of CNTs strongly inhibited the propagation of bacteria
on the filters with the antibacterial rate at each test time of 61.90% (40 minutes),
88.57% (80 minutes), and 97.86% (120 minutes), respectively.
214 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

(a) (b) (e) 100

80

Antibacterial rate (%)


60

(c) (d) 40

20

0
40 80 120
Time (min)

Figure 6.19 (a) The antibacterial results of the pristine filter for 40, 80, and 120 minutes.
(b), (c), and (d) were the antibacterial results of the composite filter for 40, 80, and 120 minutes,
respectively. (e) Antibacterial rate of the composite filter at different test time (40, 80, and
120 minutes). Source: Zhao et al. 2015 [75]. Reprinted with permission of Royal Society of
Chemistry.

6.4.8 Air Filtration and Oil Removal


In 2018, Chen and coworkers designed a multifunctional inorganic aerogels
with the combination of PM removal and oil/water separation [136]. Hydrox-
yapatite (HAP) nanowire-based inorganic aerogels were fabricated in large
quantity using HAP nanowires by freeze drying, forming three-dimensional
interconnected highly porous meshwork structure in aerogels. The as-prepared
HAP nanowire aerogel showed ultralow density (8.54 mg cm−3 ), high porosity
(∼99.7%), high elasticity, and ultralow thermal conductivity (0.0387 W m−1 K−1 ).
The as-prepared HAP nanowire aerogel can be used as the highly efficient air
filter with high PM2.5 filtration efficiency and low airflow resistance. As shown
in Figure 6.20a, the removal efficiency for PM2.5 increased with the increasing
thickness of the aerogel filter, reaching at 99% with the filter thickness of 9 mm
at high PM concentrations up to 1800 μg m−3 . The aerogel remained stable after
standing 120 hours for continuous air purification (Figure 6.20b). Additionally,
the hydrophobic HAP nanowire aerogel was applicable for continuous oil/water
separation (Figure 6.20c).
In 2018, Wang and coworkers demonstrated a superoleophobic surface that
can improve the filtration efficiency for the separation of small oil mists from the
air with low airflow resistance [137]. Such superoleophobic surface was prepared
by using perfluoroalkyl acrylic copolymer-coated commercial glass fibrous filter
via the dip-coating method. As illustrated in Figure 6.21a,b, oil droplets tended
to spread along oleophilic fibers and accumulated in the intersection region. In
contrast, when small oil mists reached the superoleophobic filters, they would
easily bounce back and forth, became larger by colliding with each other, and
eventually drained away along the fibers. As a consequence, the superoleophobic
treatment showed a significant increase in oil mist filtration. A 1.12 mm thick
superoleophobic filter showed a filtration efficiency of 99.44% for small oil mists
and almost 100% for large oil mists.
6.4 Applications 215

100 100
Removal efficiency (%)

Removal efficiency (%)


80 80

60 60 PM2.5 concentration:
3 mm, 2.51 cm s−1
3 mm, 3.35 cm s−1 300–600 (μg m–3)
40 40 Velocity: 3.35 cm s–1
6 mm, 2.51 cm s−1
6 mm, 3.35 cm s−1
20 9 mm, 2.51 cm s−1 20
9 mm, 3.35 cm s−1
0 0
0 300 600 900 1200 1500 1800 0 20 40 60 80 100 120
(a) PM2.5 concentration (μg m–3) (b) Time (h)

Oil absorption
Pump

Hydrophobic
HAP aerogel

Oil phase

Water phase

(c)

Figure 6.20 (a) PM2.5 removal efficiencies of the hydrophobic HAP nanowire aerogel filter with
different thicknesses at different PM2.5 concentrations and airflow velocities. (b) PM2.5 removal
efficiencies of the hydrophobic HAP nanowire aerogel filter (thickness 9 mm) for a long period
of time (120 hours). (c) Schematic illustration of the continuous oil/water separation device
prepared using the hydrophobic HAP nanowire aerogel. Source: Zhang et al. 2018 [136].
Reprinted with permission of American Chemical Society.

Filter
Downstream
Upstream

Re-entrainment
(a) Drainage
Bounce Collision Drainage
Superoleophobic
Fiber
2
1

1
Coalescence

Oil mist Oleophilic 2


1

(b)

Figure 6.21 (a) Oil disposal routes during oil mist filtration. (b) Schematic illustration of oil
mist interaction with fibers with different surface wettabilities. Source: Wei et al. 2018 [137].
Reprinted with permission of Royal Society of Chemistry.
216 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

6.5 Conclusion
In conclusion, nanofibrous air filters have evolved rapidly in the past 10 years
and have shown great PM2.5 removal performance along with other desirable
features (e.g. high transparency, large-scale production, high thermal stability,
thermal management, electricity assistance, the removal of toxic gases chemi-
cals, antibacterial property, and oil removal). However, there are challenges lying
ahead as summarized in the following.
In the current designs, the majority of the nanofibers were deposited on non-
woven substrates to construct composite filter media, and the active nanofibrous
layers could not stand alone due to their low mechanical strength. Thus, increas-
ing mechanical strength and the filtration performances of the nanofibrous layers
deserve more attention. Continuous and inexpensive production procedure of
nanofibrous air filters must be developed to further reduce their production cost
and affordability.
Many of the current nanofibrous filter designs have multiple functions, but
they are put together plainly without synergy. Thus one of the future directions
of nanofibrous air filters is to have more intelligence with multifunctions being
smartly integrated into one device with feedback communication cycle to maxi-
mize its performance.
The self-cleaning and/or antifouling capability would improve the filters’
longevity, and better thermal management can further increase their comfort
during their use [100]. The incorporation of energy harvesting and generating
materials (e.g. piezoelectric or triboelectric materials) would make possible some
unprecedented applications, such as air filter with self-powered environmental
sensors and air filters with their own lighting systems.
Last but not the least, the interaction mechanisms between PM particles and
nanofibers have been paid little attention in the past and are largely unclear, so
more fundamental and detailed experimental investigations are warranted. With
a clearer understanding to the interaction mechanisms, more effective air filter
can thus be rationally designed and fabricated in the future.

References
1 Liang, C.-S., Duan, F.-K., He, K.-B., and Ma, Y.-L. (2016). Review on recent
progress in observations, source identifications and countermeasures of
PM2.5 . Environment International 86: 150–170.
2 Seinfeld, J.H. (1989). Urban air pollution: state of the science. Science
243 (4892): 745–752.
3 Maricq, M.M. (2007). Chemical characterization of particulate emissions
from diesel engines: a review. Journal of Aerosol Science 38 (11): 1079–1118.
4 Vega-Lugo, A.-C. and Lim, L.-T. (2009). Controlled release of allyl isoth-
iocyanate using soy protein and poly (lactic acid) electrospun fibers. Food
Research International 42 (8): 933–940.
5 Hughes, G.J., Ryan, D.J., Mukherjea, R., and Schasteen, C.S. (2011). Protein
digestibility-corrected amino acid scores (PDCAAS) for soy protein isolates
References 217

and concentrate: criteria for evaluation. Journal of Agricultural and Food


Chemistry 59 (23): 12707–12712.
6 Dickson, J.S. and Koohmaraie, M. (1989). Cell surface charge characteristics
and their relationship to bacterial attachment to meat surfaces. Applied and
Environmental Microbiology 55 (4): 832–836.
7 Brook, R.D., Rajagopalan, S., Pope, C.A. et al. (2010). Particulate matter air
pollution and cardiovascular disease. Circulation 121 (21): 2331–2378.
8 Finlayson-Pitts, B.J. and Pitts, J.N. (1997). Tropospheric air pollution: ozone,
airborne toxics, polycyclic aromatic hydrocarbons, and particles. Science
276 (5315): 1045–1051.
9 Chow, J.C., Watson, J.G., Lowenthal, D.H. et al. (1993). PM10 and PM2.5
compositions in California’s San Joaquin Valley. Aerosol Science and Technol-
ogy 18 (2): 105–128.
10 Dockery, D.W., Schwartz, J., and Spengler, J.D. (1992). Air pollution and
daily mortality: associations with particulates and acid aerosols. Environmen-
tal Research 59 (2): 362–373.
11 Han, L., Zhou, W., Li, W., and Li, L. (2014). Impact of urbanization level on
urban air quality: a case of fine particles (PM2.5 ) in Chinese cities. Environ-
mental Pollution 194: 163–170.
12 Hu, G., Zhang, Y., Sun, J. et al. (2014). Variability, formation and acidity
of water-soluble ions in PM2.5 in Beijing based on the semi-continuous
observations. Atmospheric Research 145: 1–11.
13 Huang, B., Liu, M., Ren, Z. et al. (2013). Chemical composition, diurnal
variation and sources of PM2.5 at two industrial sites of South China. Atmo-
spheric Pollution Research 4 (3): 298–305.
14 Song, C., Pei, T., and Yao, L. (2015). Analysis of the characteristics and evo-
lution modes of PM2.5 pollution episodes in Beijing, China during 2013.
International Journal of Environmental Research and Public Health 12 (2):
1099–1111.
15 Song, Y., Tang, X., Xie, S. et al. (2007). Source apportionment of PM2.5 in
Beijing in 2004. Journal of Hazardous Materials 146 (1-2): 124–130.
16 Zheng, M., Salmon, L.G., Schauer, J.J. et al. (2005). Seasonal trends in PM2.5
source contributions in Beijing, China. Atmospheric Environment 39 (22):
3967–3976.
17 Cohen, A.J., Brauer, M., Burnett, R. et al. (2017). Estimates and 25-year
trends of the global burden of disease attributable to ambient air pollution:
an analysis of data from the Global Burden of Diseases Study 2015. The
Lancet 389 (10082): 1907–1918.
18 Liu, C., Hsu, P.-C., Lee, H.-W. et al. (2015). Transparent air filter for
high-efficiency PM2.5 capture. Nature Communications 6: 6205.
19 Pui, D.Y., Chen, S.-C., and Zuo, Z. (2014). PM2.5 in China: measurements,
sources, visibility and health effects, and mitigation. Particuology 13: 1–26.
20 Singh, K.P., Mohan, D., Tandon, G., and Gupta, G. (2002). Vapor-phase
adsorption of hexane and benzene on activated carbon fabric cloth: equi-
libria and rate studies. Industrial & Engineering Chemistry Research 41 (10):
2480–2486.
218 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

21 Dimotakis, E., Cal, M., Economy, J. et al. (1995). Chemically treated


activated carbon cloths for removal of volatile organic carbons from gas
streams: evidence for enhanced physical adsorption. Environmental Science
& Technology 29 (7): 1876–1880.
22 Han, X. and Naeher, L.P. (2006). A review of traffic-related air pollution
exposure assessment studies in the developing world. Environment Interna-
tional 32 (1): 106–120.
23 Zhang, Y., Yuan, S., Feng, X. et al. (2016). Preparation of nanofibrous
metal-organic framework filters for efficient air pollution control. Journal
of the American Chemical Society 138 (18): 5785–5788.
24 Spilak, M.P., Karottki, G.D., Kolarik, B. et al. (2014). Evaluation of building
characteristics in 27 dwellings in Denmark and the effect of using par-
ticle filtration units on PM2.5 concentrations. Building and Environment
73: 55–63.
25 Fisk, W.J. (2013). Health benefits of particle filtration. Indoor Air 23 (5):
357–368.
26 Wang, C.-s. and Otani, Y. (2012). Removal of nanoparticles from gas streams
by fibrous filters: a review. Industrial & Engineering Chemistry Research
52 (1): 5–17.
27 Wang, N., Yang, Y., Al-Deyab, S.S. et al. (2015). Ultra-light 3D nanofibre-nets
binary structured nylon 6-polyacrylonitrile membranes for efficient filtra-
tion of fine particulate matter. Journal of Materials Chemistry A 3 (47):
23946–23954.
28 Liu, J., Pui, D.Y., and Wang, J. (2011). Removal of airborne nanoparticles
by membrane coated filters. Science of the Total Environment 409 (22):
4868–4874.
29 Uppal, R., Bhat, G., Eash, C., and Akato, K. (2013). Meltblown nanofiber
media for enhanced quality factor. Fibers and Polymers 14 (4): 660–668.
30 Barhate, R.S. and Ramakrishna, S. (2007). Nanofibrous filtering media: fil-
tration problems and solutions from tiny materials. Journal of Membrane
Science 296 (1-2): 1–8.
31 Chang, J., Zhang, L., and Wang, P. (2018). Intelligent environmental nanoma-
terials. Environmental Science: Nano 5 (4): 811–836.
32 Silva, G.A., Czeisler, C., Niece, K.L. et al. (2004). Selective differentia-
tion of neural progenitor cells by high-epitope density nanofibers. Science
303 (5662): 1352–1355.
33 Gong, G., Zhou, C., Wu, J. et al. (2015). Nanofibrous adhesion: the twin of
gecko adhesion. ACS Nano 9 (4): 3721–3727.
34 Thavasi, V., Singh, G., and Ramakrishna, S. (2008). Electrospun nanofibers
in energy and environmental applications. Energy & Environmental Science
1 (2): 205–221.
35 Sahay, R., Kumar, P.S., Sridhar, R. et al. (2012). Electrospun composite
nanofibers and their multifaceted applications. Journal of Materials Chem-
istry 22 (26): 12953–12971.
36 Zhao, X., Li, Y., Hua, T. et al. (2017). Low-resistance dual-purpose air fil-
ter releasing negative ions and effectively capturing PM2.5 . ACS Applied
Materials & Interfaces 9 (13): 12054–12063.
References 219

37 Kakade, M.V., Givens, S., Gardner, K. et al. (2007). Electric field induced
orientation of polymer chains in macroscopically aligned electrospun
polymer nanofibers. Journal of the American Chemical Society 129 (10):
2777–2782.
38 Kadam, V.V., Wang, L., and Padhye, R. (2018). Electrospun nanofibre mate-
rials to filter air pollutants – a review. Journal of Industrial Textiles 47 (8):
2253–2280.
39 Zhang, R., Liu, C., Hsu, P.-C. et al. (2016). Nanofiber air filters with
high-temperature stability for efficient PM2.5 removal from the pollution
sources. Nano Letters 16 (6): 3642–3649.
40 Xu, J., Liu, C., Hsu, P.-C. et al. (2016). Roll-to-roll transfer of electrospun
nanofiber film for high-efficiency transparent air filter. Nano Letters 16 (2):
1270–1275.
41 Furukawa, H., Cordova, K.E., O’Keeffe, M., and Yaghi, O.M. (2013). The
chemistry and applications of metal-organic frameworks. Science 341 (6149):
1230444.
42 Férey, G., Millange, F., Morcrette, M. et al. (2007). Mixed-valence
Li/Fe-based metal-organic frameworks with both reversible redox and
sorption properties. Angewandte Chemie International Edition 46 (18):
3259–3263.
43 Kitagawa, S. (2014). Metal-organic frameworks (MOFs). Chemical Society
Reviews 43 (16): 5415–5418.
44 Mei, Y., Wang, Z., and Li, X. (2013). Improving filtration performance of
electrospun nanofiber mats by a bimodal method. Journal of Applied Poly-
mer Science 128 (2): 1089–1094.
45 Gao, H., Yang, Y., Akampumuza, O. et al. (2017). A low filtration resistance
three-dimensional composite membrane fabricated via free surface elec-
trospinning for effective PM2.5 capture. Environmental Science: Nano 4 (4):
864–875.
46 Zhao, X., Wang, S., Yin, X. et al. (2016). Slip-effect functional air filter for
efficient purification of PM2.5 . Scientific Reports 6: 35472.
47 Khalid, B., Bai, X., Wei, H. et al. (2017). Direct blow-spinning of nanofibers
on a window screen for highly efficient PM2.5 removal. Nano Letters 17 (2):
1140–1148.
48 Qin, X.-H. and Wang, S.-Y. (2008). Electrospun nanofibers from crosslinked
poly (vinyl alcohol) and its filtration efficiency. Journal of Applied Polymer
Science 109 (2): 951–956.
49 Choi, H.-J., Kim, S.B., Kim, S.H., and Lee, M.-H. (2014). Preparation of
electrospun polyurethane filter media and their collection mechanisms for
ultrafine particles. Journal of the Air & Waste Management Association
64 (3): 322–329.
50 Sambaer, W., Zatloukal, M., and Kimmer, D. (2012). 3D air filtration model-
ing for nanofiber based filters in the ultrafine particle size range. Chemical
Engineering Science 82: 299–311.
51 Wang, Z., Zhao, C., and Pan, Z. (2015). Porous bead-on-string poly (lactic
acid) fibrous membranes for air filtration. Journal of Colloid and Interface
Science 441: 121–129.
220 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

52 Zhang, S., Liu, H., Yin, X. et al. (2017). Tailoring mechanically robust poly
(m-phenylene isophthalamide) nanofiber/nets for ultrathin high-efficiency air
filter. Scientific Reports 7: 40550.
53 Patanaik, A., Jacobs, V., and Anandjiwala, R.D. (2010). Performance evalu-
ation of electrospun nanofibrous membrane. Journal of Membrane Science
352 (1-2): 136–142.
54 Hsiao, H.Y., Huang, C.M., Liu, Y.Y. et al. (2012). Effect of air blowing on
the morphology and nanofiber properties of blowing-assisted electrospun
polycarbonates. Journal of Applied Polymer Science 124 (6): 4904–4914.
55 Li, Q., Xu, Y., Wei, H., and Wang, X. (2016). An electrospun polycarbonate
nanofibrous membrane for high efficiency particulate matter filtration. RSC
Advances 6 (69): 65275–65281.
56 Wang, C., Wu, S., Jian, M. et al. (2016). Silk nanofibers as high efficient and
lightweight air filter. Nano Research 9 (9): 2590–2597.
57 Wang, N., Wang, X., Ding, B. et al. (2012). Tunable fabrication of
three-dimensional polyamide-66 nano-fiber/nets for high efficiency fine
particulate filtration. Journal of Materials Chemistry 22 (4): 1445–1452.
58 Souzandeh, H., Johnson, K.S., Wang, Y. et al. (2016). Soy-protein-based
nanofabrics for highly efficient and multifunctional air filtration. ACS
Applied Materials & Interfaces 8 (31): 20023–20031.
59 Souzandeh, H., Wang, Y., and Zhong, W.-H. (2016). “Green” nano-filters:
fine nanofibers of natural protein for high efficiency filtration of particulate
pollutants and toxic gases. RSC Advances 6 (107): 105948–105956.
60 Wang, S., Zhao, X., Yin, X. et al. (2016). Electret polyvinylidene fluo-
ride nanofibers hybridized by polytetrafluoroethylene nanoparticles for
high-efficiency air filtration. ACS Applied Materials & Interfaces 8 (36):
23985–23994.
61 Wang, N., Raza, A., Si, Y. et al. (2013). Tortuously structured polyvinyl chlo-
ride/polyurethane fibrous membranes for high-efficiency fine particulate
filtration. Journal of Colloid and Interface Science 398: 240–246.
62 Wang, N., Zhu, Z., Sheng, J. et al. (2014). Superamphiphobic nanofibrous
membranes for effective filtration of fine particles. Journal of Colloid and
Interface Science 428: 41–48.
63 Wang, N., Si, Y., Wang, N. et al. (2014). Multilevel structured polyacry-
lonitrile/silica nanofibrous membranes for high-performance air filtration.
Separation and Purification Technology 126: 44–51.
64 Yeom, B.Y., Shim, E., and Pourdeyhimi, B. (2010). Boehmite nanoparti-
cles incorporated electrospun nylon-6 nanofiber web for new electret filter
media. Macromolecular Research 18 (9): 884–890.
65 Cheng, Z., Zhang, Y., Han, Z. et al. (2016). A novel preparation of
anti-layered poly (vinylalcohol)-polyacrylonitrile (PVA/PAN) membrane
for air filtration by electrospinning. RSC Advances 6 (88): 85545–85550.
66 Jing, L., Shim, K., Toe, C.Y. et al. (2016). Electrospun polyacrylonitrile-ionic
liquid nanofibers for superior PM2.5 capture capacity. ACS Applied Materials
& Interfaces 8 (11): 7030–7036.
67 Wan, H., Wang, N., Yang, J. et al. (2014). Hierarchically structured polysul-
fone/titania fibrous membranes with enhanced air filtration performance.
Journal of Colloid and Interface Science 417: 18–26.
References 221

68 Zhang, S., Liu, H., Yin, X. et al. (2016). Anti-deformed polyacrylonitrile/


polysulfone composite membrane with binary structures for effective air
filtration. ACS Applied Materials & Interfaces 8 (12): 8086–8095.
69 Chuang, Y.-H., Hong, G.-B., and Chang, C.-T. (2014). Study on particulates
and volatile organic compounds removal with TiO2 nonwoven filter prepared
by electrospinning. Journal of the Air & Waste Management Association
64 (6): 738–742.
70 Vanangamudi, A., Hamzah, S., and Singh, G. (2015). Synthesis of hybrid
hydrophobic composite air filtration membranes for antibacterial activity
and chemical detoxification with high particulate filtration efficiency (PFE).
Chemical Engineering Journal 260: 801–808.
71 Uyar, T., Balan, A., Toppare, L., and Besenbacher, F. (2009). Electrospin-
ning of cyclodextrin functionalized poly (methyl methacrylate)(PMMA)
nanofibers. Polymer 50 (2): 475–480.
72 Uyar, T., Havelund, R., Nur, Y. et al. (2009). Molecular filters based on
cyclodextrin functionalized electrospun fibers. Journal of Membrane Science
332 (1-2): 129–137.
73 Cho, B.M., Nam, Y.S., Cheon, J.Y., and Park, W.H. (2015). Residual charge
and filtration efficiency of polycarbonate fibrous membranes prepared by
electrospinning. Journal of Applied Polymer Science 132 (1): 41340.
74 Wang, Z., Pan, Z., Wang, J., and Zhao, R. (2016). A novel hierarchical struc-
tured poly (lactic acid)/titania fibrous membrane with excellent antibacterial
activity and air filtration performance. Journal of Nanomaterials 2016:
6272983.
75 Zhao, Y., Zhong, Z., Low, Z.-X., and Yao, Z. (2015). A multifunctional
multi-walled carbon nanotubes/ceramic membrane composite filter for
air purification. RSC Advances 5 (112): 91951–91959.
76 Singh, V.K., Ravi, S.K., Sun, W., and Tan, S.C. (2016). Transparent nanofi-
brous mesh self-assembled from molecular LEGOs for high efficiency air
filtration with new functionalities. Small 13 (6): 1601924.
77 Zhong, Z., Xu, Z., Sheng, T. et al. (2015). Unusual air filters with ultrahigh
efficiency and antibacterial functionality enabled by ZnO nanorods. ACS
Applied Materials & Interfaces 7 (38): 21538–21544.
78 Tanaka, S., Doi, A., Nakatani, N. et al. (2009). Synthesis of
ordered mesoporous carbon films, powders, and fibers by direct
triblock-copolymer-templating method using an ethanol/water system.
Carbon 47 (11): 2688–2698.
79 Wang, D., Sun, G., and Chiou, B.S. (2007). A high-throughput, controllable,
and environmentally benign fabrication process of thermoplastic nanofibers.
Macromolecular Materials and Engineering 292 (4): 407–414.
80 Qiu, P. and Mao, C. (2010). Biomimetic branched hollow fibers templated by
self-assembled fibrous polyvinylpyrrolidone structures in aqueous solution.
ACS Nano 4 (3): 1573–1579.
81 Wang, X. and Li, Y. (2002). Selected-control hydrothermal synthesis of α-and
β-MnO2 single crystal nanowires. Journal of the American Chemical Society
124 (12): 2880–2881.
82 Fong, H., Chun, I., and Reneker, D. (1999). Beaded nanofibers formed during
electrospinning. Polymer 40 (16): 4585–4592.
222 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

83 Ding, B., Kim, H.-Y., Lee, S.-C. et al. (2002). Preparation and characterization
of nanoscaled poly (vinyl alcohol) fibers via electrospinning. Fibers and Poly-
mers 3 (2): 73–79.
84 Ding, B., Kimura, E., Sato, T. et al. (2004). Fabrication of blend biodegrad-
able nanofibrous nonwoven mats via multi-jet electrospinning. Polymer
45 (6): 1895–1902.
85 Wu, J., Wang, N., Zhao, Y., and Jiang, L. (2013). Electrospinning of multilevel
structured functional micro-/nanofibers and their applications. Journal of
Materials Chemistry A 1 (25): 7290–7305.
86 Doshi, J. and Reneker, D.H. (1995). Electrospinning process and applications
of electrospun fibers. Journal of Electrostatics 35 (2-3): 151–160.
87 Huang, Z.-M., Zhang, Y.-Z., Kotaki, M., and Ramakrishna, S. (2003).
A review on polymer nanofibers by electrospinning and their applications in
nanocomposites. Composites Science and Technology 63 (15): 2223–2253.
88 Reneker, D.H. and Chun, I. (1996). Nanometre diameter fibres of polymer,
produced by electrospinning. Nanotechnology 7 (3): 216–223.
89 Bognitzki, M., Czado, W., Frese, T. et al. (2001). Nanostructured fibers via
electrospinning. Advanced Materials 13 (1): 70–72.
90 Reneker, D.H., Yarin, A.L., Fong, H., and Koombhongse, S. (2000). Bending
instability of electrically charged liquid jets of polymer solutions in electro-
spinning. Journal of Applied Physics 87 (9): 4531–4547.
91 Chen, Y., Zhang, S., Cao, S. et al. (2017). Roll-to-roll production of
metal-organic framework coatings for particulate matter removal. Advanced
Materials 29 (15): 1606221.
92 Betha, R., Behera, S.N., and Balasubramanian, R. (2014). 2013 South-
east Asian smoke haze: fractionation of particulate-bound elements
and associated health risk. Environmental Science & Technology 48 (8):
4327–4335.
93 Zhao, S., Chen, L., Li, Y. et al. (2015). Summertime spatial variations in
atmospheric particulate matter and its chemical components in different
functional areas of Xiamen, china. Atmosphere 6 (3): 234–254.
94 Sun, Y., Wang, Z., Fu, P. et al. (2013). Aerosol composition, sources and
processes during wintertime in Beijing, China. Atmospheric Chemistry and
Physics 13 (9): 4577–4592.
95 Cheng, Z., Jiang, J., Fajardo, O. et al. (2013). Characteristics and health
impacts of particulate matter pollution in China (2001-2011). Atmospheric
Environment 65: 186–194.
96 Thorpe, A. and Harrison, R.M. (2008). Sources and properties of
non-exhaust particulate matter from road traffic: a review. Science of the
Total Environment 400 (1): 270–282.
97 Qian, Z., Wang, Z., Chen, Y. et al. (2018). Superelastic and ultralight poly-
imide aerogels as thermal insulators and particulate air filters. Journal of
Materials Chemistry A 6 (3): 828–832.
98 Liu, K., Liu, C., Hsu, P.-C. et al. (2018). Core-shell nanofibrous materials
with high particulate matter removal efficiencies and thermally triggered
flame retardant properties. ACS Central Science 4 (7): 894–898.
References 223

99 Laoutid, F., Bonnaud, L., Alexandre, M. et al. (2009). New prospects in flame
retardant polymer materials: from fundamentals to nanocomposites. Materi-
als Science & Engineering R: Reports 63 (3): 100–125.
100 Yang, A., Cai, L., Zhang, R. et al. (2017). Thermal management in
nanofiber-based face mask. Nano Letters 17 (6): 3506–3510.
101 Han, C.B., Jiang, T., Zhang, C. et al. (2015). Removal of particulate matter
emissions from a vehicle using a self-powered triboelectric filter. ACS Nano
9 (12): 12552–12561.
102 Gu, G.Q., Han, C.B., Lu, C.X. et al. (2017). Triboelectric nanogenerator
enhanced nanofiber air filters for efficient particulate matter removal. ACS
Nano 11 (6): 6211–6217.
103 Jeong, S., Cho, H., Han, S. et al. (2017). High efficiency, transparent,
reusable, and active PM2.5 filters by hierarchical Ag nanowire percolation
network. Nano Letters 17 (7): 4339–4346.
104 Bai, Y., Han, C.B., He, C. et al. (2018). Washable multilayer triboelectric air
filter for efficient particulate matter PM2.5 removal. Advanced Functional
Materials 28 (15): 1706680.
105 Zhou, Z., Dionisio, K.L., Verissimo, T.G. et al. (2013). Chemical composi-
tion and sources of particle pollution in affluent and poor neighborhoods of
Accra, Ghana. Environmental Research Letters 8 (4): 044025.
106 Khan, F.I. and Ghoshal, A.K. (2000). Removal of volatile organic compounds
from polluted air. Journal of Loss Prevention in the Process Industries 13 (6):
527–545.
107 Barea, E., Montoro, C., and Navarro, J.A. (2014). Toxic gas
removal-metal-organic frameworks for the capture and degradation of toxic
gases and vapours. Chemical Society Reviews 43 (16): 5419–5430.
108 Britt, D., Tranchemontagne, D., and Yaghi, O.M. (2008). Metal-organic
frameworks with high capacity and selectivity for harmful gases. Proceedings
of the National Academy of Sciences 105 (33): 11623–11627.
109 DeCoste, J.B. and Peterson, G.W. (2014). Metal-organic frameworks for air
purification of toxic chemicals. Chemical Reviews 114 (11): 5695–5727.
110 Wisser, D., Wisser, F.M., Raschke, S. et al. (2015). Biological chitin-MOF
composites with hierarchical pore systems for air-filtration applications.
Angewandte Chemie International Edition 54 (43): 12588–12591.
111 Hamon, L., Serre, C., Devic, T. et al. (2009). Comparative study of hydrogen
sulfide adsorption in the MIL-53 (Al, Cr, Fe), MIL-47 (V), MIL-100 (Cr), and
MIL-101 (Cr) metal-organic frameworks at room temperature. Journal of the
American Chemical Society 131 (25): 8775–8777.
112 McKinlay, A.C., Xiao, B., Wragg, D.S. et al. (2008). Exceptional behavior
over the whole adsorption-storage-delivery cycle for NO in porous metal
organic frameworks. Journal of the American Chemical Society 130 (31):
10440–10444.
113 Kumar, P., Kim, K.-H., Kwon, E.E., and Szulejko, J.E. (2016). Metal-organic
frameworks for the control and management of air quality: advances and
future direction. Journal of Materials Chemistry A 4 (2): 345–361.
114 López-Maya, E., Montoro, C., Rodríguez-Albelo, L.M. et al. (2015).
Textile/metal-organic-framework composites as self-detoxifying filters for
224 6 Emerging Nanofibrous Air Filters for PM2.5 Removal

chemical-warfare agents. Angewandte Chemie International Edition 54 (23):


6790–6794.
115 Mondloch, J.E., Katz, M.J., Isley, W.C. III et al. (2015). Destruction of chemi-
cal warfare agents using metal-organic frameworks. Nature Materials 14 (5):
512–516.
116 Koo, W.-T., Jang, J.-S., Qiao, S. et al. (2018). Hierarchical metal-organic
framework assembled membrane filter for efficient removal of particulate
matter. ACS Applied Materials & Interfaces 10 (23): 19957–19963.
117 Bian, Y., Wang, R., Wang, S. et al. (2018). Metal-organic framework-based
nanofiber filters for effective indoor air quality control. Journal of Materials
Chemistry A 6 (32): 15807–15814.
118 Dai, Z., Su, J., Zhu, X. et al. (2018). Multifunctional polyethylene
(PE)/polypropylene (PP) bicomponent fiber filter with anchored nanocrys-
talline MnO2 for effective air purification. Journal of Materials Chemistry A
6 (30): 14856–14866.
119 Rong, S., Zhang, P., Liu, F., and Yang, Y. (2018). Engineering crystal facet
of α-MnO2 nanowire for highly efficient catalytic oxidation of carcinogenic
airborne formaldehyde. ACS Catalysis 8 (4): 3435–3446.
120 Sekine, Y. (2002). Oxidative decomposition of formaldehyde by metal oxides
at room temperature. Atmospheric Environment 36 (35): 5543–5547.
121 Xiu, Z.-m., Zhang, Q.-b., Puppala, H. L., Colvin, V. L. and Alvarez, P. J.
(2012). Negligible particle-specific antibacterial activity of silver nanoparti-
cles. Nano Letters 12 (8): 4271–4275.
122 Quadros, M.E. and Marr, L.C. (2011). Silver nanoparticles and total aerosols
emitted by nanotechnology-related consumer spray products. Environmental
Science & Technology 45 (24): 10713–10719.
123 Wang, X., Yang, F., Yang, W., and Yang, X. (2007). A study on the antibac-
terial activity of one-dimensional ZnO nanowire arrays: effects of the
orientation and plane surface. Chemical Communications (42): 4419–4421.
124 Applerot, G., Lipovsky, A., Dror, R. et al. (2009). Enhanced antibacterial
activity of nanocrystalline ZnO due to increased ROS-mediated cell injury.
Advanced Functional Materials 19 (6): 842–852.
125 Ghule, K., Ghule, A.V., Chen, B.-J., and Ling, Y.-C. (2006). Preparation and
characterization of ZnO nanoparticles coated paper and its antibacterial
activity study. Green Chemistry 8 (12): 1034–1041.
126 Okyay, T.O., Bala, R.K., Nguyen, H.N. et al. (2015). Antibacterial properties
and mechanisms of toxicity of sonochemically grown ZnO nanorods. RSC
Advances 5 (4): 2568–2575.
127 Chen, X. and Mao, S.S. (2007). Titanium dioxide nanomaterials: synthe-
sis, properties, modifications, and applications. Chemical Reviews 107 (7):
2891–2959.
128 Lee, W.S., Park, Y.-S., and Cho, Y.-K. (2015). Significantly enhanced antibac-
terial activity of TiO2 nanofibers with hierarchical nanostructures and
controlled crystallinity. Analyst 140 (2): 616–622.
129 Alboofetileh, M., Rezaei, M., Hosseini, H., and Abdollahi, M. (2014). Antimi-
crobial activity of alginate/clay nanocomposite films enriched with essential
oils against three common foodborne pathogens. Food Control 36 (1): 1–7.
References 225

130 Noyce, J., Michels, H., and Keevil, C. (2006). Potential use of copper surfaces
to reduce survival of epidemic meticillin-resistant Staphylococcus aureus in
the healthcare environment. Journal of Hospital Infection 63 (3): 289–297.
131 Feng, Q.L., Wu, J., Chen, G. et al. (2000). A mechanistic study of the
antibacterial effect of silver ions on Escherichia coli and Staphylococcus
aureus. Journal of Biomedical Materials Research 52 (4): 662–668.
132 Jiang, Z., Zhang, H., Zhu, M. et al. (2018). Electrospun soy-protein-based
nanofibrous membranes for effective antimicrobial air filtration. Journal of
Applied Polymer Science 135 (8): 45766.
133 Kang, S., Herzberg, M., Rodrigues, D.F., and Elimelech, M. (2008). Antibac-
terial effects of carbon nanotubes: size does matter! Langmuir 24 (13):
6409–6413.
134 Nel, A., Xia, T., Mädler, L., and Li, N. (2006). Toxic potential of materials at
the nanolevel. Science 311 (5761): 622–627.
135 Jia, G., Wang, H., Yan, L. et al. (2005). Cytotoxicity of carbon nanomaterials:
single-wall nanotube, multi-wall nanotube, and fullerene. Environmental Sci-
ence & Technology 39 (5): 1378–1383.
136 Zhang, Y.-G., Zhu, Y.-J., Xiong, Z.-C. et al. (2018). Bioinspired ultralight
inorganic aerogel for highly efficient air filtration and oil-water separation.
ACS Applied Materials & Interfaces 10 (15): 13019–13027.
137 Wei, X., Chen, F., Wang, H. et al. (2018). Efficient removal of aerosol
oil-mists using superoleophobic filters. Journal of Materials Chemistry A
6 (3): 871–877.

You might also like