Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Journal of Aerosol Science 167 (2023) 106078

Contents lists available at ScienceDirect

Journal of Aerosol Science


journal homepage: www.elsevier.com/locate/jaerosci

A comprehensive review of particle loading models of fibrous


air filters
Gentry Berry, Ivan Beckman, Heejin Cho *
Institute for Clean Energy Technology, Mississippi State University, 205 Research Blvd, Starkville, MS, 39759, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Fibrous filters are commonly utilized to remove aerosols from the air to capture contaminants and
Fibrous filtration pollutants prior to atmospheric release, or to clean atmospheric air itself. An ideal filter achieves
Particle loading models the highest capture efficiency possible while providing the least airflow resistance possible. From
Analytical
this, the ability to predict how a filter operates based upon its functional parameters, such as
Computational
Experimental
filtration velocity and particle size and type, and characteristic parameters, such as its fiber size
and media porosity, are integral to not only properly applying and estimating filter performance,
but also facilitating the improvement and development of filtration technology. The purpose of
this work is to review both classical and contemporary particle loading models to provide a
comprehensive discussion built upon a foundation of first principles. The different relevant
particle capture mechanisms are presented and discussed in terms of fibrous air filtration, along
with the separate loading regimes. Specific loading models are then introduced considering these
concepts and are separated into three categories: 1) analytical models, 2) computational models,
and 3) experimental models. In short summary, analytical models have been well-covered and
benefit from their ease of use but suffer from a difficulty in describing the stochastic nature of
realistic particle loading. Computational models benefit from their inherent incorporation of first
principles thereby addressing the dynamic nature of particle loading, however they may be
difficult to apply outside of research environments due to their complexity and computationally
demanding nature. Experimental models benefit from the avoidance of simplifying assumptions
and the ability to correlate data describing phenomena that are difficult to quantify but are
limited to applications covered by their experimental parameters and have difficulty addressing
anomalies that present themselves. Suggested future work from identified research gaps focus on
the lack of available data and reliable methods to measure media characteristics as well as
developing models that predict the full spectrum of the loading process based upon media and
particle characteristics.

1. Introduction

An air filter may be defined as a device having a fibrous or porous medium used to remove suspended particles from the air or a gas
(ASME, 2017). Two of the primary characteristics of a filter are its filtration efficiency and resistance to airflow, where a filter ideally
operates with the highest efficiency possible while simultaneously operating with the least resistance to fluid flow possible. The ability

* Corresponding author.
E-mail address: cho@me.msstate.edu (H. Cho).

https://doi.org/10.1016/j.jaerosci.2022.106078
Received 8 April 2022; Received in revised form 10 September 2022; Accepted 14 September 2022
Available online 24 September 2022
0021-8502/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

to accurately predict the pressure drop of a filter as a function of particle mass loading over a period time has been sought for many
decades. Yet a conclusive and generally accepted method remains elusive, although many possible candidates have been proposed
which offer promising results. The purpose of this study is to provide a review of important mechanisms and concepts related to filter
loading as well as many of the proposed models for describing this process and its evolution over time. Previous work reviewing this
subject has offered excellent insight, such as the work published by Thomas (Thomas, 2017) covering analytical and experimental
work on the filtration of solid aerosols, or the work of Abdolghader et al. (Abdolghader, Brochot, Haghighat, & Bahloul, 2018) which
focuses specifically on nanoparticle filtration. What differentiates the current work is the consideration of the contributions made from
computational models and the comparison of them to analytical and empirical models, while also considering work covering the full
spectrum of particle sizes that may be reasonably expected.
Before discussing the concepts related to filter loading, it is beneficial to first examine why a comprehensive model describing its
evolution is worth achieving. Simply put, filtration models allow for the prediction of filter performance and the duration of a filter’s
service life. This implies that more accurate filtration models will allow for greater confidence in the performance of the filter in
sensitive applications, such as nuclear air filtration. The ASME AG-1 Code on Nuclear Air and Gas Treatment (AG-1) (ASME, 2017)
defines a maximum pressure drop for a clean filter that is allowable for use; however, specifications and criteria for an unclean, loaded
filter are largely neglected. The Department of Energy’s (DOE) Nuclear Air Cleaning Handbook (Angle et al., 2003) suggests that a
facility replaces a used filter once the pressure differential exceeds a specified limit. Further, it is also recommended that the filters are
discarded once a prescribed amount of time has been exceeded, implying a definite shelf life. This is due to an assumed degradation of
the structural integrity of the filter media over time. Thus, the ability to predict the performance of a filter given the expected operating
conditions offers a method to mitigate the substantial cost associated with their use, which is several times greater than their initial
cost. For example, not only are the High Efficiency Particulate Air (HEPA) filters used in nuclear facilities very expensive, but they also
require testing to verify performance criteria before they are able to be used, as well as a substantial cost associated with their disposal.
Therefore, the total cost of each filter is potentially many times greater than the cost of initial purchase. Exerting more control over the
purchased inventory will inevitably allow for a reduction in wasted filters as it will offer greater insight into the functional lifetime of
the filter, therefore providing a more accurate estimation of the number of required filters over a given period of time. Heating,
Ventilation, and Air Conditioning (HVAC) systems offer a simpler example of the benefit of comprehensive and accurate filtration
models. Liu et al. (Liu, Claridge, & Deng, 2003) state that the need for accurate models stems from the significant amount of consumed
energy in air handling units from pressure loss due to filters, illustrated in Fig. 1, and that simple and accurate analytical models and
algorithms are necessary for energy modelling. Jung (Jung, 1987) studied the effect of the condition of air filters, and their effects on
the performance of HVAC systems and the life of its equipment, noting that air filters could assist in preventing heat exchanger fouling,
and may offer cost effective solutions to extend the life of the HVAC equipment.
While is it not possible to write about all the numerous applications of air filtration, another aspect worth mentioning is the removal
of aerosols from the environment and its effect on human health. In the last several years, the COVID-19 pandemic has renewed interest
in the control and mitigation of airborne droplets carrying the virus indoors and through HVAC systems. Although many technologies
exist that could address these airborne droplets, common fibrous air filters offer a well-known and effective method to reduce the
number of viral particles in the air, thereby reducing the probability of infection in an individual (Berry, Parsons, Morgan, Rickert, &
Cho, 2022). Particulate Matter (PM) in an environment has also been linked with many detrimental health effects. Kim et al. (Kim,
Kabir, & Kabir, 2015) state that the presence of PM poses a greater risk to human health than ground-level ozone or other common
pollutants, such as carbon monoxide, and that effective air quality management is necessary to reduce that risk. Mukherjee and
Agrawal (Mukherjee & Agrawal, 2017) note a negative relationship between exposure of people to inhalable particles equal to or
smaller than 10 μm (PM10) and their health. Agay-Shay et al. (Agay-Shay et al., 2013) studied gestational exposure to air pollution and
concluded that exposure of the mother to increased concentrations of PM10 was associated with multiple congenital heart defects.
Coronas et al. (Coronas et al., 2009) investigated people living near an oil refinery in Brazil and noted that the results suggested that the
PM10 from the refinery increased level of DNA damage in lymphocytes. In an alternate manner to other research, Keuken et al.
(Keuken, Zandveld, van den Elshout, Janssen, & Hoek, 2011) estimated that an additional 13 months of life were added to individuals

Fig. 1. Pressure loss versus airflow rate (Liu et al., 2003).

2
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

after a decrease in PM10 and other atmospheric aerosols. Fibrous air filters play an important role in ensuring the cleanliness of the air
in our indoor environments, given their prevalence as an affordable air cleaning technology. From this, it may be implied that an
accurate and comprehensive filtration model would not only assist in the making of more effective filters, but would also enable the
creation of engineered solutions, where a filter may be manufactured as a tailored solution for a specific pollutant that may be
prevalent in a specific area.
The above examples only cover very specific uses of air filters; however, they effectively illustrate situations that would greatly
benefit from comprehensive loading models. More generally, a comprehensive model will facilitate more informed decisions based
upon the life cycle and expected performance of the filters and will assist in understanding the implications of the physical charac­
teristics of the media from a manufacturing standpoint.

2. Conceptual basis

2.1. Particle deposition mechanisms and capture efficiencies

Particles are deposited onto filter fibers through five particle deposition mechanisms (Hinds, 1999): 1) interception, 2) inertial
impaction, 3) diffusion, 4) gravitational settling, and 5) electrostatic attraction. However, in many cases only the interception, inertial
impaction, and diffusion mechanisms are analyzed, as typical fibrous air filters do not utilize charged fibers to employ electrostatic
attraction, and the gravitational settling may be neglected over the timescale appropriate for fibrous filtration. These are loosely
illustrated in Fig. 2, with the aerosol following streamlines from left to right. According to Lee and Ramamurthi (Lee & Ramamurthi,
2001), interception is defined as the deposition of a particle which follows a streamline that brings the particle center to within one
radius from the surface of a fiber. Interception efficiency may be approximated using the Kuwabara flow field as shown in Equation (1),
where α represents the void fraction, also called the solidity, Ku represents the Kuwabara hydrodynamic factor, and R represents the
interception parameter.

1 − α R2
ηinter = (1)
Ku (1 + R)
Inertial impaction occurs when a particle breaks free from the streamline it is following due to its inertia, contacts, and subse­
quently sticks to a filter fiber. Stechkina et al. (Stechkina, Kirsch, & Fuchs, 1969) derived an expression for the inertial impaction
efficiency using the Kuwabara flow field, shown in Equation (2), where St represents the Stokes number for a fiber. The equation is
valid over 0 ≤ R ≤ 0.4 and 0.0035 ≤ α ≤ 0.111.
St [( ) ]
ηimp = 2
29.6 − 28α0.62 R2 − 27.5R2.8 (2)
(2Ku)
According to Hinds (Hinds, 1999), diffusion is a direct result of the Brownian motion of particles and is characterized by a

Fig. 2. Common deposition mechanisms.

3
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

seemingly random trajectory caused by collisions with gas molecules. Lee and Liu proposed an expression for the deposition efficiency
from diffusion, again using the Kuwabara flow field shown by Equation (3), where Pe represents the Peclet number.
( )1
1− α 3 −
(3)
2
ηdiff = 2.58 Pe 3
Ku
It is worth mentioning that the classical theory of filtration, i.e. the single-fiber theory, considers all the deposition mechanisms as
acting independently of each other. For example, particles collected by interception would not interfere or enhance the particle
collection by impaction. Although this is realistically incorrect, it may be considered as approximate, especially at the onset of
filtration prior to a significant amount of deposition. In reality, a particle or group of particles that deposit onto a fiber through
impaction, for example, may also begin collecting particles through interception, thus enhancing the overall collection efficiency of the
fiber. However, although these interactions are known, they are difficult to quantify and express analytically. Inasmuch, the enhanced
interception of particles due to diffusion may be given by Equation (4) (Hinds, 1999):
2
1.24R3
ηdr = 1 (4)
(Ku ⋅ Pe)2
Hinds (Hinds, 1999) also gives an expression for the efficiency due to gravitational settling as shown in Equation (5), where VTS
represents the settling velocity for a particle and U0 represents the velocity of the gas.
VTS VTS
ηgrav = (1 + R) or − (1 + R) (5)
U0 U0
The expression may be positive or negative depending on whether the particle is traveling in the same direction as its natural
( )2
settling motion, or in the opposite direction. If the particle motion is horizontal, the efficiency is on the order of VUTS0 .Finally, for a
comprehensive view of the classical theory, it is important to mention particle capture by electric forces (where the efficiency is
denoted by the variable ηelec ). Namely, this is deposition due to either the particles, fibers, or both having an electrical charge, but it is
usually neglected unless the charge on the fibers or particles is somehow known.The reader is referred to the work of Brown (Brown,
1993) for a more thorough discussion on particle deposition due to electric forces.
Each efficiency given above constitutes a single mechanism for particle collection by a fiber that are happening simultaneously in
reality. Given the specific conditions that the filter media is subjected to, not all the deposition mechanics will be effective, and some
may be completely neglected for simplification. The single-fiber efficiency, being the ratio of the number of captured particles by a
fiber to the total number of particles based upon its projected area normal to the gas flow, is predicated upon the assumption that each
of the deposition mechanics is functioning independently of the other deposition mechanics. Yet the sum of the independent mech­
anisms cannot exceed 100%, as shown by Equation (6) (Hinds, 1999). Thus, for example, the particles that are captured by inertial
impaction do not affect the particles that are captured by interception (Lee & Ramamurthi, 2001). Although this is not the case in
reality, this greatly simplifies the analysis, and does not yield egregious errors in many common circumstances. Given the classical
approach, the total single-fiber efficiency is given by Equation (6) (Hinds, 1999).
( )( ) ( )
ηΣ = 1 − (1 − ηinter ) 1 − ηimp 1 − ηdiff (1 − ηdr ) 1 − ηgrav (1 − ηelec ) (6)

Equation (6) may be further simplified when particle capture mechanisms are not relevant. An example of this is the work of Lee
and Liu (Lee & Liu, 1980, 1982) who presented previously derived expressions for the single fiber efficiency model in the region of the
Most Penetrating Particle Size (MPPS) based upon the Kuwabara flow field and found good agreement between their theory and
experimental data. In the region of the MPPS, it was stated that particle capture by interception and diffusion were considered to be
dominant and were summed together for the total single fiber efficiency as shown in Equation (7). Capture by diffusion decreases with
increasing particle size, while capture by interception increases. Thus, one mechanism will be dominant while the other remains less
relevant in the region of MPPS, which is consistent with Equation (6) operating under the assumption that each capture mechanism
acts independently.
η = ηdiff + ηinter (7)
The final form of the single fiber efficiency equation proposed by Lee and Liu is given in Equation (8), which accounted for realistic
fibrous media by correlating experimental data and adding empirical constants.
( )13 /
1− α 1 − α R2
(8)
/
23
η = 1.6 Pe− + 0.6
Ku Ku 1 + R
However, further alterations to this equation were proposed by Payet et al. (Payet, Boulaud, Madelaine, & Renoux, 1992) to ac­
count for slip when the size of the fibers approaches the mean free path of air, as well as to account for very fine particles at low Peclet
numbers. The modified single-fiber efficiency equation of Payet et al. is given below in Equation (9).
( ) 2
1 − α 13 ’’ − 23 1 − α R’
(9)
/ /
η = 1.6 Pe + 0.6
Ku Ku 1 + R’

4
G. Berry et al.
5

Journal of Aerosol Science 167 (2023) 106078


Fig. 3. Scan of typical deposit structures. A) Longitudinal view of particle deposits with cross-sectional views given with the section ranges in μm. B) A composite image of all the cross-sectional view
from part A) (Kasper et al., 2010).
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

In Equation (9), Pe′′ and R are the modified Peclet number and modified Interception Parameter, respectively, given below in

Equations (10) and (11).


⎛ ⎞− /
32
( )⎜ ⎟
⎜ 1 ⎟
(10)
/
Pe′′ = Pe ⋅ Cd− 32 ⎜ ⎟
⎝1 + 1.6( 1− α)13 Pe− Cd ⎠
/ /
23
Ku

′2
R R2
= Cr (11)
1 + R′ 1 + R
The correction factor for the efficiency by diffusion, Cd , and the correction factor for the efficiency by interception, Cr , are given
below in Equations (12) and (13) where Knf represents the Knudsen Number relative to the fiber.
( )13 /
(1 − α)Pe
Cd = 1 + 0.388 ⋅ Knf (12)
Ku

1.999 ⋅ Knf
Cr = 1 + (13)
R
It is worth noting that when Lee and Liu’s efficiency model was used to calculate the MPPS (Lee & Liu, 1980), a substantial dif­
ference was recorded between the model and experimental results at higher velocities. This was attributed to the model neglecting the
inertial impaction capture mechanism, which is more relevant at higher velocities.
Considering the different available expressions estimating fiber capture efficiency, the single-fiber efficiency may now be related to
the total efficiency for filtration media as shown by Equation (14).
( )
4
η = 1 − P = 1 − exp − ηΣ α t (14)
πdf
From an inspection of the single-fiber efficiency and its constituent components, it is clear that several factors will have a profound
impact on the performance of the filter. For example, the interception parameter may be regarded as relevant based upon either the
diameter of the fiber or the diameter of the particle. Thus, deposition by interception may be enhanced by either increasing the size of
the particles, or by decreasing the size of the fibers. However, an alternative line of thought may be considered where particles of
different diameters within a polydisperse distribution are collected at different efficiencies. Deposition by impaction is slightly more
complex, as it is influenced by the Kuwabara hydrodynamic factor, interception parameter, and solidity. But unlike interception,
impaction is also influenced by the Stokes number, accounting for the particle’s inertia. Thus, particles are not only collected based
upon their diameter size, but also their velocity.

2.2. Microscale Particle loading

Now that a conceptual basis for particle collection has been established through the single-fiber theory, it is worth noting that the
single-fiber efficiency that has been discussed up to now has been the efficiency of a bare fiber, with no prior particle collection. As
particles are collected, they eventually deposit onto other particles forming chain-like agglomerates or more compact structures
(Davies, 1973; Kasper, Schollmeier, & Meyer, 2010). An illustration for this is given in Fig. 3. For solid particles, the van der Waals
force is mainly responsible for adhesion to fibers or other particles. Furthermore, although the detachment and resuspension of
previously adhered particles is a realistic possibility, models typically assume this does not occur and that once a particle adheres to a
fiber, it remains stationary. Realistic particle collection is further complicated by idea of particle bounce, which clearly decreases the
probability of particle collection, although it is worth noting that a particle will be subjected to many opportunities through filtration
media of sufficient thickness to be captured. The effect of particle bounce on single-fiber efficiency is discussed by Kasper et al. (Kasper,
Schollmeier, Meyer, & Hoferer, 2009) who highlighted previous work relating the concept of the probability of particle adhesion to the
single-fiber collection efficiency as shown in Equation (15).
η=φ·h (15)
In this equation, φ represents the collision efficiency, and h represents the adhesion probability. A reasonable assumption may be
made for submicron particles and relatively modest filtration velocities that h = 1. Dahneke (Dahneke, 1971) proposed that the
possibility exists that a particle will bounce based upon the particle’s coefficient of restitution and the energy interaction between the
particle and the surface. This interaction between the particle and the surface is conceptualized as a potential well, where the particle
will escape capture by the fiber through particle bounce if its velocity is above a certain threshold. Brach and Dunn (Brach & Dunn,
1998) proposed that for an oblique collision between a particle and a fiber that a significant amount of kinetic energy may remain,
resulting in a sliding or rolling motion in the particle on the fiber. The particle may either deposit elsewhere on the fiber, such as further
along the sides instead of directly where contact was made, or may become resuspended. Paw U (Paw U, 1984) introduces the concept
of a critical Stokes number, which is derived from a critical rebound speed where the particles begin to bounce upon contact with a

6
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

fiber. It is interesting that in this work the collection efficiency was noted at first to increase with an increasing Stokes number, but then
decrease sharply once the critical Stokes number was reached. Dunn and Renken (Dunn & Renken, 1987) also studied the effects of the
Stokes number, Reynolds number, and the interception parameter on the collection efficiency of particles on a single wire. Although
the particles used in the study were relatively large (0.254 μm–25.4 μm), it was shown that the collection efficiency was lower than
what was predicted by theory. This finding was related again to the particles’ incident velocity, where the two critical velocities were
presented. The first critical velocity indicated a point where particles at a lower velocity will never reach the surface of the fiber, and
the second critical velocity indicated a point at which the particles at a greater velocity will always rebound from the surface and
escape capture.
Continuing the research into particle bounce, Kasper et al. (Kasper et al., 2010) discussed the idea that the structure of collected
particles will change based upon the filtration parameters, i.e., the Stokes number, Peclet number, and interception parameter that
describe particle collection. A model was presented from correlated data for a single, isolated fiber relating the porosity of the particle
deposit to these parameters, which is useful in understanding the growth and formation of the deposits. A representation of the
relationship between the particle deposit geometry and filtration parameters is illustrated effectively in Fig. 4. It may be seen that
collection dominated by diffusion is described by deposits around the entire fiber, collection dominated by impaction is deposited
preferentially on the upstream side of the fiber, and the interception parameter influences the compactness of the deposit. Kasper et al.
also describe the idea of a variable called the bounce parameter based upon the ratio between the Stokes number and interception
parameter. From this, a critical bounce parameter describes the conditions for the onset of particle bounce. Particles with a bounce
parameter greater than the critical value will bounce and either escape capture or will deposit onto the fiber and create compact,
forward-facing structures. However, particles with a bounce parameter less than the critical value will tend to form dendritic structures
with an appreciable branching to the sides. This trend is also present within Fig. 4. The authors suggest in this study that the filtration
parameters and subsequent particle deposits significantly affect fiber collection efficiency, as is expected, resulting in deviations from
the single-fiber theory as more particles are captured by a fiber. The importance of this conclusion is encountered in subsequent
sections, as many analytical, computational, and empirical models rely on the single-fiber theory and its derivative forms in some
fashion.
On the subject of nanoparticle filtration specifically, Wang and Otani (Wang & Otani, 2013) studied the effects of particle shape on
the collection efficiency and particle capture mechanisms. They did not present a new model, however their results led to some
interesting conclusions. Wang and Otani observed that differently shaped particles settled onto fibers in different ways. For example, a
square shaped particle would tend to slide on the fiber immediately after contact, whereas a spherical particle tended to roll, thus
causing an overall difference between the final location on the fiber and whether the particle detached from the fiber. This is

Fig. 4. The relationship between the particle deposit and filtration parameters. Reproduced by Kasper et al. (Kasper et al., 2010) from Kanaoka et al.
(Kanaoka, Emi, Hiragi, & Myojo, 1986).

7
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

noteworthy for empirical work, as test particles employed in lab environments may take different shapes, yielding slightly different
results for the same filtration parameters.
Further complicating the concepts of particle collection is the fact that filtration media is comprised of many closely packed fibers
interacting with each other. Thus, neighboring fibers will have a pronounced effect on the particle deposition to other fibers through
their impact on fluid flow. Early ways to account for these effects were given by popular cell models, two of which are the Happel and
Kuwabara models (Happel, 1959; Kuwabara, 1959), which differ in an assumption related to concentric boundaries around the fibers
and their conditions, allowing for a solution for the fluid flow and streamlines. Although the Happel and Kuwabara cell models are
popular and widespread in use, other models similarly account for the effects of neighboring fibers, such as the work presented by
Spielman and Goren (Spielman & Goren, 1968). Instead of defining cells around each fiber used in their computational model, the
authors used a model that incorporates damping effects from neighboring fibers into the flow around each fiber. Emi et al. (Emi,
Okuyama, & Adachi, 1977) calculated the collection efficiency numerically using a cell model and compared it to experimental data.
The authors found that the collection efficiency from particle interception and inertial impaction increased for increasing solidities.
Furthermore, they determined a theoretical difference in the collection efficiencies between an isolated fiber and a fiber with neighbors
which illustrates the functional difference between calculations using the single-fiber theory and calculations accounting for the effects
of neighboring fibers. Yeh and Liu (Yeh & Liu, 1974a; 1974b) studied the effects of realistic fiber geometry and the effect of particle
collection by impaction, interception, and diffusion simultaneously by numerically solving a modified Kuwabara-Happel flow field.
These results were coupled with experiments to validate the theoretical predictions and showed good agreement. From this study, the
authors asserted that the treatment of the particle collection mechanisms as independent is arbitrary, conceptually inaccurate, and
adds explicit error from neglecting its simultaneous nature.

2.3. Macroscale Particle loading

From the preceding discussion on particle bounce and particle capture, the question arises as to how particles will deposit onto a
filter fiber subjected to continuous loading. Payatakes and Gradon (Payatakes & Gradoń, 1980) presented a theoretical model focusing
on the formation of dendritic structures created by the capture of solid particles. Although only focusing on inertial impaction and
interception, it was shown that the dendrites grew independently for a time, but eventually showed interference with the other
dendrites, which would entangle and grow together as a single structure. This concept of particle deposit growth and entanglement
forms the basis of the transition particle loading regime from the depth of the media to the surface of the media into the particle cake.
Furthermore, an interesting concept is presented by the authors, who suggest that analytical solutions to the problems presented by
filtration mechanics is exceedingly difficult due to the random nature of the processes. However, it was proposed that a stochastic
method of simulation is more feasible than an analytical method based upon implicitly incorporating mechanics such as the random
nature of dendrite growth, with its main drawback being that the solution is not analytical in form and may present other difficulties,
such as studying Brownian diffusion deposition.
In its simplest form, the loading of filter media is divided into three regimes: 1) depth loading, 2) surface loading, and 3) a transition
period between depth loading and surface loading. The increase in pressure drop for depth loading is approximately linear, as is the
overall increase in pressure drop for the surface loading. However, the transition period between depth loading and surface loading is
non-linear. An exact definition for the end of the depth loading regime and the beginning of the surface loading regime does not exist,
although several authors have proposed logical arguments in an attempt to define such a point. Japuntich (Japuntich, 1991) simply
describes this point as the “upper transition point,” where the pressure drop curve becomes linear again after the transition period. It
was assumed that this point must correspond to the formation of the particle cake, since the slope of the line was the same using the
same monodisperse aerosol, regardless of the media type that was challenged. Alternatively, Bourrous et al. (Bourrous et al., 2014)
proposed that a linear correlation coefficient R2 could be used with the deviation from a perfect correlation (1 – R2) to describe the

Fig. 5. General loading curve for flat sheet media.

8
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

clogging point and thus the transition to surface filtration. Fig. 5 illustrates an example loading curve as would be expected in a generic
situation. A transition point between the different regimes would effectively allow for an estimation between depth and surface loading
which in turn allows for the utilization of different regime models in sync.

3. Analytical loading models

The discussion of the analytical models begins with an introduction to fundamental equations describing the pressure drop through
fibrous media and packed beds of spheres, and a summary of the included references may be seen in Table 1. Traditionally, an
empirical correlation proposed by Davies (Davies, 1953), has been used to describe the pressure drop through fibrous media, while the
Kozeny-Carman equation and other derivative forms such as the Ergun equation (Ergun & Orning, 1949) have been used to describe
the pressure drop through the particle cake. In his pioneering work, Davies developed a dimensionless group based upon Darcy’s law of
flow through porous media, shown below in Equation (16).

4ΔPAr2 ( )
= 64α1.5
f 1 + 56α3f (16)
μQt
In this equation, ΔP represents the pressure drop, A the filtration area, r the fiber radius, μ the air viscosity, Q the volumetric air
flow rate, t the height or thickness of the filter media, and αf the packing density of the fibrous media. The original work proposed by
Kozeny and improved by Carman (Tien & Ramarao, 2013), shown below in Equation (17), describes the pressure drop of a creeping
fluid flow through a packed bed of solids.
( )2
μtv 1 − αp
ΔP = 36Kk (17)
dp2 α3p

Here, t represents the thickness of the particle cake, αp represents the packing density of the particle bed, dp represent the diameter of
the particles, and Kk is the Kozeny constant, which is usually approximated to be equal to 5. The work proposed by Ergun yields similar
results but is based off of empirical constants for experiments conducted using a number of gases and porous mediums. This equation is
shown in Equation (18), where the first and second terms account for the viscous and inertial effects of the flow, respectfully.
( )2 ( )
μtv 1 − αp ρtv2 1 − αp
ΔP = 150 + 1.75 (18)
dp2 α3p dp α3p

Abdolghader et al. (Abdolghader et al., 2018) suggested that clogging models may be separated into two categories: particulate
models and capillary models, and noted that the Kozeny-Carman equation is a capillary model. However, capillary models were cited
to be applicable only to sufficiently small packing densities, and therefore the particulate models were more suited for nanoparticle
deposits, as these tended to have large packing densities. Thus, the Kozeny-Carman equation may not always be the correct choice for
particle cake modeling, although it is the certainly one of the more commonplace models. The effects of humidity on hygroscopic and

Table 1
Notable analytical air filtration loading models.
Author Details

Davies (Davies, 1953) Empirical correlation of pressure drop model


Ergun and Orning (Ergun & Orning, 1949) Creeping fluid flow through packed bed with empirical correlation
Tien and Romarao (Tien & Ramarao, 2013) Pressure drop analyzed as creeping fluid flow through packed bed of
solids
Abdolghader et al. (Abdolghader et al., 2018) Review article of fibrous filtration of nanoparticles
Bergman et al. (Bergman, 2006; Bergman, Taylor, Miller, Biermann, & Hebard, 1979, p. First substantial particle fiber loading model
43, 1981, 1983)
Letourneau et al. (Letourneau, Mulcey, & Vendel, 1990, pp. 800–812, 1992) Investigated effects of particle distribution inside depth of filter
media
Thomas et al. (Thomas et al., 1999, 2001) Modified Bergman model by layering depth loading
Bemer and Calle (Bémer & Callé, 2000) Pressure drop and efficiency over stationary and dynamic regimes
Hinds and Kadrichu (Hinds & Kadrichu, 1997) Characterization of dust loading via particle fiber model
Dhaniyala and Liu (Dhaniyala & Liu, 2001b) Incorporates inhomogeneities with local packing densities
Shou et al. (Shou, Fan, Zhang, Qian, & Ye, 2015) Investigated randomness of fiber orientation
Schweers and Loffler (Schweers & Löffler, 1994) Structural inhomogeneities of fibrous filter material
Podgorski (Podgórski, 2009) Study of fiber size distributions in filter media
Payatakes and Okuyama (Payatakes & Okuyama, 1982) Behavior of homogeneous filters with differential equations
Kanoaka and Hiragi (Kanaoka & Hiragi, 1990) Prediction of pressure drop of dust-loaded filter
Payatakes (Payatakes, 1976, 1977) Theoretical behavior of filter during dynamic loading
Song and Park (Song & Park, 2006) Analytical solution with empirical correlation
Laborde et al. (Laborde, Fabbro, Mocho, & Ricciardi, 2002) Modeled clogging for pleated filters
Hu et al. (Hu et al., 2019, pp. 207–210) Examined influence of fiber size and porosity using a machine
learning model
Xiao et al. (Xiao et al., 2019) Derivation of a fractal model of Kozeny-Carman constant

9
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

non-hygroscopic particles were also investigated. It was stated that a high relative humidity could possibly increase the capillary force
between fibers and particles, but that it was only significant for larger particles.

3.1. Depth loading models

The first loading model in our discussion was proposed by Bergman et al. (Bergman, 2006; Bergman et al., 1979, 1981, 1983), and
is prevalent due to its relative simplicity and applicability to HEPA filter media. Although mostly applied to HEPA media, it may be
applied to other, less efficient media as long as the degree of error from the simplifying assumptions and correlations is deemed to be
acceptable. Bergman proposed a generalized loading model that accounts for the behavior of the loading curve throughout the life
cycle of the filtration media, beginning in the depth loading regime and ending in the surface loading regimes. This loading model is, at
its core, a combination between the Kozeny-Carman equation and the single-fiber model as proposed by Davies using Darcy’s law. As
particles are deposited onto the individual fibers, dendrites form and begin to become preferential deposition sites for new particles
due to the concepts discussed above. These dendrites are then modeled as newly formed fibers with a diameter equal to the mean
diameter of the deposited particles and are considered to behave identically to the original filter fibers. Relating the volume fraction of
the particles to the deposited mass allows for the prediction of the pressure drop as a function of deposited mass, shown in Equation
(19), where m represents deposited mass, ρpd represents the density of the particles within the deposit, and A represents the effective
filtration area.
m
αp = (19)
ρpd At
Accounting for the clean media pressure drop yields the complete and simplified version of the model, given in Equation (20), and
incorporates Davies’ correlation for depth loading and the Kozeny-Carman equation for surface loading.

Fig. 6. Pressure drop versus surface mass for different filtration velocities (Letourneau et al., 1990, pp. 800–812).

10
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

m
ΔP − ΔPo = θv (20)
dpn ρpd A

√̅̅̅̅
α2p
where, θ = 64μ for depth loading, θ = 180μ (1− for surface loading, n = 1 for depth loading, and n = 2 for surface loading
αf
df αp )3
characterized by a packed bed. It is worth noting here that it is important for the amount of depth loading to be small, as Bergman’s
model does not account for the spatial effects of realistic filtration, meaning that particles and their resulting structures are assumed to
be homogeneously distributed throughout the entire depth of the media. Although this is simply not the case for realistic filtration, the
difference becomes irrelevant for high efficiency media as the majority of particles will deposit close to the surface, and the transition
from depth loading to surface loading begins quickly. Thus, the simplifying assumption is acceptable for high efficiency media as the
differential loading level within the depth of the filter media will not have a pronounced effect on the performance of the media as a
whole. It is worth noting that Bergman addressed criticism of his model (Bergman, 2006), suggesting that for the range of velocities
normally encountered in HEPA filtration, the depth loading was inconsequential and could be ignored as the majority of mass would be
deposited at a shallow depth and on the surface of the media.
Considering the practicality of applying the Bergman model, it is important to note its limitation regarding performance prediction,
namely due to a lack of knowledge of the physical characteristics of the particle cake during surface loading. Currently and to the
knowledge of the authors, no generally accepted method outside of a laboratory exists to estimate either the volume fraction or the
thickness of the particle cake as a function of the deposited particles. Furthermore, the Bergman model relies on other variables that are
difficult to reliably obtain for the filtration media, such as the porosity of the media itself, which suffers from the same difficulties
mentioned above. Without either of these prescribed unknowns, it is not possible to complete the calculation reliably as a prediction,
and ultimately its use may be limited to theoretical explanations of gathered data. However, if deemed acceptable then the appropriate
assumption may be made using traditionally assumed values, such as the typical porosity of HEPA media. Some expressions exist that

Fig. 7. Distribution of deposition depth at different filtration velocities (Letourneau et al., 1990, pp. 800–812).

11
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

attempt to describe the porosity of the particle cake, although these do not appear to be prevalent throughout published literature.
These will be discussed in more detail in later sections.
Building off of the work of Bergman et al., Letourneau et al. (Letourneau et al., 1990, 1992) investigated the effects of particle
distribution inside the depth of the filter media. The premise was that the particles would deposit with an exponential decrease as the
depth of the media increased, and that the resulting effect on the filter performance would be appreciable. This is in contrast to the
previously discussed assumption adopted in the Bergman model. The authors characterized this distribution of particle loading
throughout the depth of the media by using a so called “peeling method,” whereby both the particle and fibers were removed in layers
from a loaded flat sheet using an adhesive tape. From the completed experiments, it was shown that as the face velocity increased, the
percent of particles deposited into the depth of the media decreased. However, it is also worth noting that the vast majority of the
particles were deposited at a shallow depth regardless of the face velocity, illustrated in Figs. 6 and 7. Furthermore, it was suggested
that the amount of aerosol penetrating into the depth of the filter media was insignificant once a benchmark of deposited mass per unit
area had been reached (2.5 g/cm3 in the experiments conducted by the authors). The reasoning behind this was that the loading
mechanism had transitioned from depth loading to surface loading.
To account for the loading distribution of the particles in the depth of the filter media, the Bergman model was modified to use a
volume fraction corresponding to the particles that was a function of the depth of the medium, shown in Equation (21).

m ζe− ζx
αp (x) = (21)
A ρ(1 − e− ζt )

Here, x represents the depth into the medium, t represents the thickness of the medium, and ζ represents a mean penetration factor of
the medium. Using the modified Bergman model in conjunction with experimental data, the penetration factor was determined, which
alleviated the necessity of continually defining the penetration profile via the peeling method, although this would then rely on the
assumption that the profile retains its exponential nature. Finally, it must be noted that the authors used relatively small particles for
their experiments, which would effectively increase the depth loading, thereby exacerbating the penetration and utility of the pro­
posed model in contrast to the original Bergman model. However, this is not to imply that the model proposed by Letourneau et al. is
not useful, as Bergman focused primarily on HEPA media, whereas the penetration of less efficient medias could potentially deviate for
the inherent assumptions in the Bergman model.
Thomas et al. (Thomas et al., 1999, 2001) furthered the work from Letourneau et al. by using another modified form of the Bergman
model, accounting for the exponential nature of the particle loading by dividing the depth of the media into layers. Thus, this modified
version potentially increased the accuracy of the model by explicitly avoiding the assumption that the particle deposits in the depth of
the media are homogeneous throughout, and instead attempts to model the top layers of the media as more loaded and therefore more
efficient. Furthermore, Thomas et al. conducted experiments to evaluate the results of the modified loading model, which showed good
agreement. It should be noted that the authors used the same particles the aforementioned work by Letourneau et al. used, which
would inherently result in a similar outcome from their experiments. However, the discretization of the media depth should inherently
account for the penetration profile not implicit with the Bergman model, regardless of particle type, by the calculation of the layer
efficiencies and their specific evolution.
The modified version of the Bergman loading model by is given below in Equation (22).
( )12 (
/
)
αf αpJ,t αf αp,J,t ( ( )3 )
ΔPJ,t = 64μvZJ + + 1 + 56 αf + αpJ,t (22)
df2 dpf2 ,J,t df dpf ,J,t

where J denotes the layer number, Z represents the thickness of each layer, and t accounts for time. The pressure drop across the filter
media is given as the sum of the pressure drop across each layer, shown below in Equation (23), where np represents the number of
layers the media has been separated into for the calculation.
np

ΔPt = ΔPJ,t (23)
J=0

Bemer and Calle (Bémer & Callé, 2000) conducted tests on two different kinds of HEPA media to determine the influence of particle
size on the evolution of filter efficiency and pressure drop. In their work, they divided the loading model into two different regimes: the
stationary regime and the dynamic regime. The stationary regime describes the phase of loading where the performance of the filter is
not changing as particles are being loaded, i.e., a clean piece of filter media that remains unaffected in its performance as a few
particles are deposited. The dynamic regime describes the phase of filter loading, that changes as particles are loaded, i.e., an increase
in pressure drop or efficiency of the filtration media due to the particle deposits captured by the fibers. It is worth noting that the
authors critiqued the use of monodisperse aerosols commonly used in a laboratory setting to describe the realistic performance of
filters, as typical aerosols found outside of a lab setting usually have a polydisperse distribution. The authors used the previously
discussed work of Lee and Liu (Lee & Liu, 1980) to calculate the particle penetration while the media was in the stationary loading
regime. This estimates the effects from the distribution of fiber diameters by accounting for the elementary packing fractions, αi , in
Equation (24), and then by calculating the total particle penetration using those elemental packing fractions and their associated
efficiencies, ηi , in Equation (25). However, the use of the Lee and Liu’s efficiency model only accounts for the diffusion and interception
collection mechanisms, so the adoption of this model limits the application of the proposed model to relatively small particles where

12
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

the impaction mechanism is not dominant.


dfi+1

( ) ( )
αi = α f df , σ f dLog df (24)
dfi

( )

n ∑
n
αi ηi
P= Pi = exp − A (25)
i=1 i=1
(1 − αi )df i

For the dynamic regime, the authors utilized the Mass Median Diameter (MMD) of a polydisperse aerosol instead of a singular
particle diameter and calculated the elemental packing fraction and pressure drop using Equations (26) and (27).
dp i+1

( ) ( ) ( )
αpi = αp f MMD, σf · η dp · dLog dp (26)
dp i

( )12 (
/
)

i
αf αpi αf αpi
ΔP = 64μtv 2
+ 2
+ (27)
i=1 df dpfi df dpfi

Experimental data was shown to have good agreement with the proposed model during depth loading with media consisting of a
homogeneous structure. However, the data quickly diverged once surface loading began, as anticipated by an understanding of the
Bergman model without its transition from the depth loading form to the surface loading form. Also, filter media with a heterogeneous
structure did not show good agreement between the measured and calculated efficiencies. These differences were attributed to an error
in the estimation of the initial fractional efficiency due to strong heterogeneity in the packing fraction and thickness of the filtration
media. It was also noted that the model underestimated the efficiency of submicron particles.
Hinds and Kadrichu (Hinds & Kadrichu, 1997) also discussed the characterization of dust loading for particles of three different
mass mean aerodynamic diameters as a function of particle size and face velocity, and analyzed penetration as a function of the same.
They presented a predictive model that simulates deposited particles as newly formed fibers. The results are consistent with other
experiments and analytical models: on a particle number basis, larger particles have more of an effect than smaller particles. However,
on a mass basis, smaller particles have a larger effect for equivalent mass loading than larger particles. This is consistent for both the
flow resistance and particle penetration. It is also worth noting that the authors pointed out that particle loading affects two parameters
directly: the average fiber diameter and the filter solidity. These both make sense, as the newly deposited particles are treated as new
fibers, and the filter solidity will evolve as more mass is deposited, i.e., more volume is occupied. Like the Bergman model, this model
also assumes that the captured particles are evenly distributed throughout the depth of the media, which will not cause issues with
efficiency medias, but will potentially lead to inaccuracies in less efficient medias. Furthermore, the model makes its calculations based
upon the total length of media fibers and particle fibers. For a clean filter, this will make relatively accurate estimates, however as
dendrites form the complexity of the collection increases, implying a point where the proposed model would no longer be capable of
accurately estimating performance.
The work discussed in this section offers relatively simple models that primarily describe the deposition of particles in the depth of
filtration media and utilize the Kozeny-Carman equation or a derivative form to address the surface loading due to the formation of the
particle cake. They are dependent on equivalent variables, such as the equivalent fiber diameter instead of a distribution of fiber
diameters like is commonly found in fibrous media. These assumptions tend to lead to views of the media being homogeneous and
uniform through its depth, which may or may not be impactful depending on the application.

3.2. Models incorporating media inhomogeneities

Dhaniyala and Liu (Dhaniyala & Liu, 2001b) incorporated filter media inhomogeneities into a model for a more realistic repre­
sentation. The collection efficiencies for each cell were calculated using the Lee and Liu efficiency model, and the total media efficiency
was calculated from the cell efficiencies, local average flowrate through the cells, and the assumed packing fraction of the media from
the incorporated distribution. Again, it is noted that the incorporation of the Lee and Liu model limits the direct applicability of this
proposed model to relatively small particles where the impaction mechanism is not dominant. However, future work could use these
models as a template and incorporate other collection mechanisms that may be relevant as well. The authors compared the results of
the analytical calculations to collected experimental data and noted a good agreement between the collected data and the calculations
over a wide range of conditions. It is worth noting that the packing fraction only varied in two dimensions and was assumed ho­
mogeneous throughout the depth of the media, whereas the work described above (Letourneau, Thomas, Bemer and Calle) assumed
that the media varied in two dimensions as a function of particle loading into the depth of the media but was homogeneous across each
of the layers. Perhaps the most interesting implication of this work is that the size of the penetrating particles was a function of the
properties of each cell, and therefore the size distribution of particles that penetrated the filter was based upon the local cell char­
acteristics instead of the overall filter media characteristics.
An alternative approach to considering the impact of media inhomogeneity is by considering the geometry of the fibers themselves.
Shou et al. (Shou et al., 2015) accomplished this by using a Voronoi diagram defined by a gamma distribution to embed fibers into the

13
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

modeled media. A notable result from this study was that the modeled media reflecting the randomness of the fibers was more
permeable than media modeled with ordered array. Furthermore, it is interesting to note that the authors reported a decrease in the
single-fiber collection efficiency as the fiber volume fraction increased. This result suggests that the effects of randomness in fiber
orientations become more prevalent within denser fibrous materials, while also suggesting that variations in the flow field are less
pronounced for looser fibrous materials. The model presented by the authors showed good agreement with experimental data by Lee
and Liu (Lee & Liu, 1981) and numerical simulations by Hosseini and Tafreshi (Hosseini & Tafreshi, 2010), although the presented
model fit closer to the numerical simulation than the theoretical model. However, it is worth noting that the work cited by Lee and Liu
regarding the single fiber efficiency is based upon an ordered array model, whereas the numerical simulation by Hosseini and Tafreshi
accounts for disordered fiber geometry.
Schweers and Loffler (Schweers & Löffler, 1994) proposed a model that reflected the structural inhomogeneities of a typical fibrous
filter by dividing a media sample into slices, allowing for evaluation of the local properties of the individual pieces. The results from
this investigation yielded several conclusions: 1) the local packing densities may be considered to be normally distributed, 2) the
distribution is a function of the mean packing density and the fiber diameter-specific element sizes, and 3) a relationship exists relating
the standard deviation of the local packing density or the respective standard deviations of the fiber length to the cubic element size
and fiber diameter. Further, it was stated that although the fibers tended to be oriented orthogonally to the flow through the media,
significant deviation existed and increased for an increasing packing density. The local flow through the medium was described as
being normally distributed with a 30% standard deviation relative to the average flow rate. For their proposed model, the local
permeability was verified to be a function of the local packing density and the angle between the fibers face flow direction and the fiber
axis. The permeability of the individual elements used in the model was calculated using the Kuwabara cell model, where it is
interesting to note that it increased with a decreasing angle between the fibers and flow direction. It is noteworthy that the authors also
incorporated particle adhesion into their model, where an adhesion fraction is introduced that modifies the analyzed collection
mechanisms to account for particle bounce. Furthermore, the authors stated that the inhomogeneities of the media impacted the
collection mechanisms differently, based upon the observation that smaller particles decreased in collection efficiency more than
larger particles. In general, the pressure drop and particle collection were both stated to decrease with an increasing amount of in­
homogeneity. A possible explanation for this is that since the pressure drop must be equivalent across each cell, an increase in in­
homogeneity resulted in the more permeable cells experiencing more flow, resulting in an overall decrease in pressure drop. Likewise,
an increase in flow would imply an increase in filtration velocity, which would directly influence the particle collection and be
especially impactful for the diffusion mechanism. Overall, the model agreed well with experimental data, and the authors proposed
that non-selective corrections to the single-fiber efficiency were not universally valid based upon the effect of media inhomogeneity.
Specifically, this statement targeted corrections to single-fiber efficiencies that are based upon empirical data alone.
The work of Podgorski and Podgorski et al. (Podgórski, 2009; Podgorski, Maißer, Szymanski, Jackiewicz, & Gradon, 2011) illus­
trated the effects of modeling media with polydisperse fiber diameters as opposed to a mean fiber diameter specifically regarding the
filtration of nanoparticles. Experimental work included the challenging of melt-blown glass media with nanoparticles and comparing
the results to classical theories based upon mean and equivalent fiber diameters. Fig. 8 shows a comparison between the penetration
results for the proposed Fully Segregated Flow Model (FSFM) and the penetration results using a geometric mean fiber diameter.
Podgorski’s model, which used a distribution of random fiber diameters predicted lower filtration efficiency and higher particle
penetration than analytical models using a mean fiber diameter. Furthermore, it was illustrated that no representation of a mean fiber
diameter, such as geometric mean or root-mean-square, or even an equivalent fiber diameter calculated from clean media pressure

Fig. 8. Relationship between penetration results of the Podgorski FSFM model for different geometric mean fiber diameters versus a monodisperse
penetration evaluated using a single equivalent diameter.

14
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

drop adequately agreed with the results for the polydisperse diameter distribution. This illustrates that simplifying assumptions
common to many air filtration models result in errors due to a deviation from realistic media geometry. It should be emphasized that
this work only utilized nanoparticles, which could potentially skew the results and conclusions if this model were to be applied to
larger particles, although there is no specific reason why this model could not be further developed for larger particles. Continuing the
work of the FSFM, the authors proposed the Partially Segregated Flow Model (PSFM), where the fibers are considered to be mixed in
the media at a state where the gas velocity is not evenly distributed due to the distribution of the geometry, but the different fiber
diameters are not completely segregated. The results of their analysis again indicated that no single mean or equivalent fiber diameter
could adequately fit the experimental data they compared it to, although different representation of the diameters did yield some
success over limited ranges. However, it is worth noting that the authors found significantly better results when an equivalent diameter
was reduced from experimental pressure drop data than when the mean diameter was used to represent a polydisperse fibrous media.

3.3. Differential equation-based models

Although traditionally most of the analytical models have been based upon Davies’ empirical model and Darcy’s law, and sub­
sequently other models such as the Bergman model and Lee and Liu’s efficiency model, it is important to recognize alternative based
models. Payatakes and Okuyama (Payatakes & Okuyama, 1982) describe the behavior of homogeneous fibrous filters with a set of
three phenomenological differential equations and corresponding boundary conditions. In this, the effects of the deposited particles
were analyzed with respect to the efficiency and other variables such as the packing density, Stokes number, and interception
parameter. Correction functions were calculated and applied to the local deposition of particles and the pressure drop. An illustration
of the results may be seen in Fig. 9, illustrating the relationship between the single-fiber efficiency and other related variables. As with
the other theoretical models, this model is only applicable in the absence of a particle cake, i.e., during the depth loading regime, and
for dendritic growth that is independent of other dendrites. However, like the Thomas model, this model utilizes a multilayer scheme
and thus should be able to describe the onset of cake formation by analysis of the local layer solidity and making an assumption about a
maximum solidity before the onset of cake formation.
Kanoaka and Hiragi (Kanaoka & Hiragi, 1990) proposed a model based upon a partial differential equation to predict the pressure
drop of a dust loaded filter. However, this model may be difficult to implement due to the solution of the partial differential equation.
Furthermore, it was stated that the dendrites resulting from diffusion and interception would have larger effective fiber diameters for
an equivalent accumulated particle volume than dendrites resulting from inertial impaction and interception. This is noteworthy as the
pressure drop is directly related to the effective fiber diameter, and thus by extension the dominant capture mechanisms. Furthermore,

Fig. 9. Single-fiber efficiency (η0) plotted again the interception parameter for different Stokes numbers (NSt) and fiber packing densities (γ). Left:
γ = 0.01 and 0.03. Right: γ = 0.05 and 0.07 (Payatakes & Okuyama, 1982).

15
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

many models do not account for dendrites forming differently by different collection mechanisms, and simply assume they may be
described by a representative particle diameter. Payatakes (Payatakes, 1976, 1977) proposed an analytical model for the dynamic
behavior of a fibrous filter during particle loading based upon a set of differential equations, although the work excludes deposition by
diffusion. Payatakes also proposed a generalized analytical model based upon the preceding work to describe the growth of particle
dendrites in a similar fashion to the Bergman model, except that it accounts for dendrite growth as a function of time and particles
deposited over a layer. An important aspect of Payatakes’ model was the concept of adhesion efficiency, which as was discussed above
is usually assumed to be equal to 100%. Furthermore, the model focused exclusively on deposition by interception and suggested how
the particles deposit and how particle structures will dynamically change and grow. Specifically, the model suggests which part of the
fiber dendrites will deposit on, which may be useful for the analysis of particle deposit concepts such as shadowing. The author
critiqued the work as it only accounts from growth due to deposition by interception, which may be contrasted with the previously
cited work referring to the impact of the particle deposit geometry based upon the dominant collection mechanisms.
Song and Park (Song & Park, 2006) developed an analytical model for the particle penetration efficiency of a polydisperse aerosol
for both diffusion and inertial impaction dominant capture mechanisms. Their work builds off of the efficiency model given by Lee and
Liu, however an empirical correlation is utilized to incorporate the effects of inertial impaction on the capture efficiency. Thus, this
model is constrained in a similar fashion to previous models by using a correlation, although its use is extended to larger particle sizes
by incorporation of the impaction mechanism. The authors stated that the presented model closely approximated an exact solution for
a monodisperse aerosol. However, when a polydisperse aerosol was analyzed, the proposed model deviated from the exact solution due
to the polydispersity bringing a relatively large number of smaller particles that penetrate the filter.

3.4. Models incorporating pleat geometry

Up to this point, most of the presented work has been focused on flat sheet media. However, flat sheet media models simplify
particle loading with an assumption that the surface of the filter media is experiencing uniform loading and does not change. Pleated
media, however, does not experience uniform loading across the entire surface of the filter media and affects the flow field entering the
media. This is discussed in more detail in later sections. Laborde et al. (Laborde et al., 2002) discussed models of clogging for pleated
filters, observing that most loading models have been developed for flat filters, whereas most industrial filters are pleated. Unlike flat
sheet media, the authors proposed that pleated media experiences an additional surface loading regime, whereby the effective
filtration area is reduced as particles are deposited in between the pleats. Through experiments the authors asserted that the rela­
tionship between the ratio of pressure drop of a loaded filter to the pressure drop of a clean filter increased for a decreasing particle size
for the same mass loading. This is consistent with results for flat sheet media as well, as smaller particles result in more surface area for
the same mass loading than larger particles. It was also shown that larger filtration velocities will result in particles with a higher
inertia, and therefore will promote particle deposition further within the pleat than lower filtration velocities as illustrated in Fig. 10.
Furthermore, with a polydisperse particle size distribution, the different inertias of the particles result in a heterogeneous deposition of
the particles within the pleat, especially at lower filtration velocities, further increasing the aforementioned pressure drop ratio.
However, by definition, the pressure drop across the entire filter face must be equivalent, thus leading to the conclusion that a dynamic
situation is present regarding the exact location that particles of different sizes may be expected to deposit. The authors concluded by
presenting two models: one derived analytically, and the other derived from previous experimental work. These models focused on the
influence of particle size, the height and depth of the pleats, and the filtration velocity. The authors showed mixed results when
comparing the data from their experiments to their analytical models, which implies factors that are not considered within their model.

3.5. Alternative models

Hu et al. (Hu et al., 2019, pp. 207–210) examined the parameters affecting the performance of filtration materials focusing on the
influence of fiber size and porosity of the filter media. This work applied a machine learning model to review and reconsider the
parameters affecting filtration. Hu et al. examined the definitions and formulas for many variables and conducted a simulation to
compare different formula combinations and variables, such as porosity and fiber size. The simulations were then compared to
experimental results; however the paper lacks a description of their experiments and no explanation was offered to address the

Fig. 10. Different possible particle loading configurations from filtration velocity. Reproduced from Laborde et al. (Laborde et al., 2002).

16
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

differences between experimental and simulated results.


Xiao et al. (Xiao et al., 2019) presented a derivation for a fractal model of the Kozeny-Carman constant in order to more closely
model a realistic, porous, fibrous media. The authors provide an interesting perspective, as it is unusual for the Kozeny-Carman
equation to be applied to the fibrous media itself since the media is typically regarded as comprised of fibers instead of the chan­
nels formed by them. This model focused on the relative roughness, porosity, area fractal dimension, and tortuosity fractal dimension
as the variables in the model. Overall, the authors cited good agreement with published experiments, and noted that the
Kozeny-Carman constant for roughened surfaces decreased with roughness, porosity, area fractal dimensions of the pores, and the
tortuosity fractal dimension. The authors also noted that the dimensionless permeability of the porous, fibrous media increased with
increasing porosity.

4. Computational models

Computational models of air filtration have made significant advancements as a compliment to analytical modeling in recent
decades. Offering a much faster and cheaper alternative to experimental methods, they can offer insight into difficult questions, such as
how the effects of neighboring fibers influence particle capture. Similar to analytical models, great care must be taken with as­
sumptions in order to ensure accuracy and avoid errors. As mentioned in the previous sections, analytical models commonly use the
assumption that the filtration media is flat, allowing solutions by assuming particles load onto the entire filtration area uniformly while
the area itself remains constant as a function of time or particle loading. However, this simplification is often unrealistic when
considering pleated filter materials, where the pleat geometry, such as pitch, depth of the pleats, angle of the pleats, etc., makes it
difficult to model analytically and expensive to investigate experimentally. Using pleated material as an example, computational
models offer a viable solution to investigate this difficult topic, specifically the influence of the pleat geometry, in an inexpensive way
that allows for the flexibility to investigate many different configurations. Table 2 summarizes twenty-two notable computational
models for air filtration discussed in detail below.

4.1. Computational models for pleat geometry

The prevalence of pleated media in the normal applications of fibrous air filters is enough to warrant investigation into its per­
formance. Fotovati et al. (Fotovati et al., 2012) developed a model in an attempt to offer a computationally affordable method of
simulating the pressure drop and collection efficiency of pleated filter media in order to isolate the effects of the specific pleat ge­
ometry. The author’s computational pleat loading model demonstrated that adding more pleats to the media resulted in a slower rate
of flow resistance increase from particle loading and was validated by a comparison to the work conducted by Thomas et al. (Thomas,
Penicot, Contal, Leclerc, & Vendel, 2001), which showed good agreement. This is a reasonable and expected outcome, as more pleats

Table 2
Notable computational air filtration models.
Author Details

Fotovati et al. (Fotovati, Tafreshi, & Pourdeyhimi, 2012) Filtration efficiency and flow resistance of pleated filter media
Fotovati et al. (Fotovati, Hosseini, Vahedi Tafreshi, & Pourdeyhimi, Filtration efficiency and flow resistance of pleated filter media
2011)
Theron et al. (Théron, Joubert, & Le Coq, 2017) Investigation of air velocity and particle capture in pleated filter media
Saleh et al. (Saleh, Fotovati, Vahedi Tafreshi, & Pourdeyhimi, Investigation of polydisperse particle loading in pleated filter media
2014)
Feng et al. (Feng, Pan, Wang, & Long, 2018) Hybrid Eulerian-Markov method for particle collection in pleated filter media
Rao and Faghri (Rao & Faghri, 1988) Comparison of ordered fiber arrays with analytical models
Fotovati et al. (Fotovati, Vahedi Tafreshi, & Pourdeyhimi, 2010) Comparison of fiber orientations, in-plane and through-plane
Tahir & Tafreshi (Tahir & Tafreshi, 2009) Study of fiber in-plane and through-plane orientations
Beckman et al. (Beckman, Berry, Cho, & Riveros, 2021) Describes a method of random fiber geometry generation
Hosseini & Tafreshi (Hosseini & Tafreshi, 2010) Comparison of ordered and disordered 2-D and 3-D models
Yang et al. (Yang et al., 2018) Study of fiber geometry with particle loading and its representation by different distribution
moments
Deshpande et al. (Deshpande, Antonyuk, & Iliev, 2020) Study of non-spherical particle cake development
Shu et al. (Shu, Qian, Zhu, & Lu, 2020) Evaluation of the performance of using OpenFOAM for fibrous filtration
Math2Market (GeoDict, 2019; Math2Market, n.d.) Comprehensive GeoDict software suite for air filtration simulation
Linden et al. (Linden, Cheng, & Wiegmann, 2018) Media replication using imaging techniques and GeoDict
Azimian et al. (Azimian, Cheng, & Wiegmann, 2020) Calculating filter permeability and flow resistance using GeoDict
Bai et al. (Bai et al., 2020) Developed a single-fiber efficiency model based upon simulated fibrous media
Maddineni et al. (Maddineni, Das, & Damodaran, 2018) Investigated particle capture and adhesion of fibrous media
Alilou (Alilou, 2018) Investigated the velocity fields and simulated the growth of particle deposits in pleated
media
Hoch et al. (Hoch, Azimian, Baumann, Behringer, & Niessner, Comparison of voxel-based and mesh-based models
2020)
Becker et al. (Becker, Cheng, Kronsbein, & Wiegmann, 2016) Investigated particle cake performance for resolved and unresolved models
Lee et al. (Lee, Jung, Park, Kim, & Kim, 2021) Multiscale approach to simulate and extrapolate microscale performance to full-scale
models

17
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

implies a larger filtration area and less particulate mass per unit area for a fixed particle concentration and flow rate. However,
Fotovati’s model also produced a less obvious result. It was shown by the authors that the flow field upstream of the media was affected
and altered as more particles were loaded onto the filter, thus illustrating the need to account for this variation. This alteration of the
flow field is objectively similar to the filtration velocity changing through cells of inhomogeneously modeled media from the previous
section. Furthermore, this result illustrates a common unrealistic idealization of many analytical models that assume uniform flow
across the entire face of the filter. This is illustrated below in Fig. 11. Thus, a distinct advantage of computational modeling is illus­
trated as the ability to analyze the changing flow field over time as more mass is loaded dynamically onto the media. A third outcome of
Fotovati’s computational model was the impact of collection efficiency as a function of pleat counts. By adding more pleats to the
media, the average face velocity through the filter decreased which resulted in a higher collection efficiency of the diffusion capture
mechanism and a lower collection efficiency by inertial impaction mechanism. This model showed the pronounced effect of pleat
geometry on particle collection efficiencies, which in turn changed the MPPS for a polydisperse particle distribution. Adding more
pleats tended to increase the MPPS, while fewer pleats tended to decrease the MPPS.
Fotovati et al. (Fotovati et al., 2011) presented a second computational model to study the filtration efficiency and flow resistance
of pleated filter media. The authors suggested that semi-analytical models primarily based on flat sheet loading were not useful in the
design of pleated media filters. Their work describes computational modeling techniques for pleated geometry, an algorithm for
particle tracking and deposition modelling, filter performance for the different conditions, and the effects of particle size and flow
velocity. Their computational model suggests that an optimal pleat count will yield minimum flow resistance regardless of the pleat
orientation or geometry. This model reiterated the reduction in flow resistance with the addition of more pleats from their earlier work.
However, their results also indicate that a higher pleat count results in higher velocities in the pleat channels and non-uniform particle
deposition on the pleats. The particle cake tended to form deeper within the pleats as a result, especially if larger particles were
considered, although this was dependent on the pleat geometry. The authors concluded by noting the effects of particle cake
permeability and total number of pleats on the changing rate of flow resistance through the filter as measured by the pressure drop
across the filter. Theron et al. (Théron et al., 2017) also produced a computational model to investigate the local velocity variations of
different pleat geometries and the overall effects on flow resistance, then validated their computational model with experimental
results. Similar to the previously cited work above, the authors related the pleat geometry to velocity gradients within the pleats, which
directly affected the particle capture mechanisms. Furthermore, it was again shown that a higher pleat pitch corresponded to more
significant velocity gradients. However, this work gave special attention to the curved portion of the pleats, as the bending of the media
induces local changes in the media properties such as an increase of solidity at the bend, implying that this detail may be impactful for
an accurate model. The authors concluded that tall and tightly packed pleats create more significant velocity gradients and a higher
face velocity, while short and wide pleats with larger pleat pitch result in smaller velocity gradients. Thus, this study corroborates the
conclusion that the pleat geometry of the filter media has a significant impact on the capture mechanisms and collection efficiency for a
clean filter. A noteworthy implication from the results of two previously cited studies is that the macroscale properties of the filter
media, i.e., the pleat geometry, will influence the microscale performance. For example, more pleats will decrease the filtration ve­
locity in the pleat channel, thus influencing the capture mechanisms. This could lead to potentially engineered solutions for pleated
filters, where not only the media is engineered on the microscale via fiber diameter size and distribution for optimal performance of
specific particle types and sizes, but the pleat geometry as well.
Saleh et al. (Saleh et al., 2014) proposed a fast mathematical model to simulate the design and performance of pleated filters prior
to manufacturing. An example of the resulting dust deposition pattern is illustrated below in Fig. 12. Their work presents several
different expressions used to calculate the permeability of the filter media and utilizes clusters of particles instead of individual
particles in order to reduce the computational time. It is interesting to note that the authors suggested calibrating the models with
experimental data prior to use. However, noting that their work was macroscale, they also suggested that microscale models may be
sufficient for calibration in the absence of experimental data. The authors’ results indicated that by adding more pleats, the flow
resistance decreased while the filtration efficiency increased at a lower rate, as expected. Furthermore, it is worthwhile noting an
observation made by the authors, where the curves for pressure drop and efficiency plotted against mass per unit area for different

Fig. 11. Contour plots of cell solidity and deposited particles, and streamlines simulating particle deposition inside of a filter pleat (Fotovati
et al., 2012).

18
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

Fig. 12. 3-D dust deposition pattern for different pleat pitches and particle sizes. (a), (b), (c), and (d) contain monodisperse particle sizes, while (e)
and (f) illustrate polydisperse particle sizes. The color bars illustrate particle cake solidity (Saleh et al., 2014)

pleat counts should collapse on each other if the flow patterns were the same for the different cases. This is to mean that the different
curves should superimpose onto each other if normalized by the appropriate variables. However, it was also noted that this condition
may only be the case for clean media at the beginning of loading. The authors also investigated the effects of loading from a poly­
disperse aerosol, specifically compared to the loading from equivalent smaller and larger monodisperse cases. It was stated that the
flow resistance from a polydisperse aerosol increased faster than from monodisperse aerosols. This behavior was attributed to the
combined effects of larger and smaller particles, namely the increased drag from the greater surface area per unit mass of the smaller
particles combined with the accelerated particle cake formation induced by the larger particles. The interplay between these two
factors results in a particle cake that forms earlier and yields an increased pressure drop per unit of particulate mass. However, this
study did not mention the effects of larger particles forming a particle cake deeper withing the pleat, although it was observed that
decreasing the particle size or increasing the number of pleats led to a decrease in particle cake uniformity. The authors suggested that
the advantage of their computational model lies in the complexity of the flow fields and filtration mechanics, where the temporal
effects of particulate loading filtration is too unwieldy for analytical solutions. Thus, this work provides a good example of how
computational models enable specific conclusions to be drawn from slight adjustments of filter geometry and flow parameters, well
beyond the capability of pure analytical models.
Feng et al. (Feng et al., 2018) proposed a hybrid Eulerian-Markov method as opposed to a more traditional Lagrangian method to
simulate the particle motion and particle collection in a pleated filter medium. This model was applied to flat-sheet geometry and
validated with published experimental data and semi-empirical models for collection efficiency as a function of loaded particle mass in
order to establish confidence in the results for the simulated pleated geometry, similar to the suggestion in the previously discussed
work regarding model calibration (Saleh et al., 2014). For the case of the flat sheet media, the proposed model yielded reasonable
results, and thus was considered accurate for more complex geometries. The authors concluded that the hybrid Eulerian-Markov
method was valid based upon the agreement with other established methods and noted a reduction in computational time between
80% and 90% when compared with a more traditional Lagrangian method. Thus, this study exemplified a functional validation useful
for the development of future work, while also showcasing a method to drastically reduce the computational time, thereby offering a
more approachable solution for CFD models limited by resources. However, this model was also based upon a capture efficiency model

19
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

to estimate the number of captured particles for each cell. Thus, it is reasonable to assume that the accuracy of this CFD model would be
limited by the accuracy of the efficiency model used. This work used the single-fiber efficiency model, which has been discussed in
detail in the previous sections.

4.2. Computational models for fiber orientation and size

While analytical models have typically been constrained by assumptions of ordered fiber geometry, cross sectional shape, cur­
vature, equivalent diameter sizes, and ordered fiber geometry, numerical models benefit from the freedom to generate realistic fiber
geometry and characteristics. The work of Rao and Faghri (Rao & Faghri, 1988) seems to be foundational for subsequently published
work regarding the computational modelling of fibrous filtration. The authors developed a numerical technique that was not restricted
to creeping flow, thus allowing for a simulation of filtration in the laminar flow regime. To do this, fibers were modeled as an ordered
array of in-line, parallel cylinders, similar to many of the analytical models. Particle capture was simulated using the interception and
diffusion capture mechanisms. The authors’ model showed good agreement with the work of Kuwabara (Kuwabara, 1959), Happel
(Happel, 1959), Hasimoto (Hasimoto, 1959), and Henry and Ariman (Henry & Ariman, 1983). However, their model predicted a
higher pressure drop than that of the experimental work conducted by Davies (Davies, 1953). A reasonable explanation for this de­
viation is that Davies’ experimental model incorporates inhomogeneities inherent to actual filtration media, whereas the authors’
work does not. The results of this work support obvious results, such as the reliance of filtration efficiency on particle size, fluid
properties, and media solidity. However, the results also suggest particle diffusion can be classified into two regimes based on the
Peclet number. For Pe > 100, diffusion is primarily a boundary layer phenomenon proportional to Pe2/3 , while for Pe < 50 the single
fiber efficiency deviates significantly from the Pe2/3 correlation. This difference was cited as likely being from the cross-stream second
derivative in the proposed model.
Fotovati et al. (Fotovati et al., 2010) studied fiber orientation and its effect on the flow resistance and collection efficiency. In this
work, the fibers themselves were aligned with different in-plane and through-plane angles to isolate the effects stemming from the
different orientations. However, this also meant that the fibers did not have a random orientation distribution and implies that the
geometries are not completely representative of realistic fibrous media. The geometry also contains fiber interpenetrations, which
would naturally lead to a deviation in the accuracy of the results, although it is difficult to know how severe that deviation would be.
Finally, the size of the fibers was large, which separates this work from much of the other work utilizing much smaller fibers. The
resulting filtration efficiency of larger particles was highest for parallel fiber with in-plane orientations, while the efficiency decreased
with an increasing angle of orientation. Flow resistance and the efficiency of smaller particles was not significantly affected by the
in-plane fiber orientation. The through-plane orientation, where fibers are no longer oriented orthogonal to the flow, resulted in a
decrease in the collection efficiency. However, it was shown that an increase in the through-plane orientation increased the quality
factor for smaller particles (nanoparticles) but that larger particles decreased the quality factor. The quality factor may be defined as a
relationship between the particle penetration and the pressure drop of the media, indicating that either less penetration or a smaller
pressure drop results in superior filtration performance. This model is useful for the analysis of clean filter media and understanding
the beginning of stages depth loading but cannot be applied to more advanced stages of loading since the collected particles were
assumed to disappear from the computational domain.
Tahir and Tafreshi (Tahir & Tafreshi, 2009) also studied the effects of fiber orientation on the permeability of the fibrous media.
This work also allowed for the interpenetration of the fibers, although a check was performed to ensure that the resulting solidity was
correct. Following the earlier analytical work of Spielman and Goren (Spielman & Goren, 1968) the authors noted the in-plane
orientation of the fibers did not have a significant effect on the permeability of the media. However, ordered arrangements of fi­
bers, such as fibers orthogonal to air flow and unidirectional fibers (fibers with the same orientation) showed a significant effect on the
permeability. Furthermore, it was also suggested that media permeability increased with fiber through-plane orientation, and further
increased as the standard deviation of the through-plane angles increased, considering that an angular distribution may be applied.
Altogether, this work suggested that the permeability of fibrous media that is highly ordered and aligned should be expected to be less
than moderately aligned or isotropic media. Regarding the description of realistic fibrous media, Beckman et al. (Beckman et al., 2021)
utilized a Python script to generate media characterized by random fiber geometry. This was accomplished by modeling the digital
fibers sequentially beginning with the bottom-most fiber and laying subsequent fibers above it until the top-most fiber is reached. Each
fiber is characterized by a flexibility, allowing for the bending of fibers over other fibers. Through this, realistic fiber geometry
characteristics such as solidity were achieved without allowing for fiber interpenetration. However, it should be noted that it is
possible to create unrealistic fibrous geometry by selecting unreasonable parameters such as the flexibility, that a typical glass fiber
would be incapable of achieving. This would potentially create errors similar to the fibrous interpenetrations previously mentioned
inasmuch as the CFD model may accept this geometry and yield results but would ultimately be incapable of accurately describing
realistic media performance.
The work of Hosseini and Tafreshi (Hosseini & Tafreshi, 2010) compared ordered and disordered 2-D and 3-D filtration models with
semi-analytical models and analyzed the effects of aerodynamic slip with regard to particle collection efficiency and pressure drop. It
was noted that ordered 2-D models tend to over-predict the pressure drop compared to more realistic 3-D disordered geometries,
whereas disordered 2-D geometries tend to predict somewhere in between the results of ordered 2-D and disordered 3-D geometries.
The inclusion of aerodynamic slip resulted in significantly reduced simulated pressure drop, which is an important result as the authors
claimed that no empirical or analytical expression accurately predicts the pressure drop for slip flow in fibrous media. Regarding the
collection efficiency, the results indicated that the disordered 2-D models tended to slightly under-predict the collection efficiency as

20
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

compared to realistic disordered 3-D geometries at low Peclet numbers, whereas the disordered 2-D models slightly over-predicted the
collection efficiency at high Peclet numbers. It should be noted that collection by inertial impaction was not included in this study.
When compared to empirically modified analytical cell models, the disordered 2-D and 3-D computational models generally had good
agreement, thus implying that the proposed computational models may be used to significantly reduce the cost and time of empirical
studies. However, when compared to empirical models found in literature that included inertial impaction, it was shown that the
difference in the single fiber efficiencies were appreciable for cases of low, medium, and high Stokes numbers alike. The use of
computational modelling to further empirical and analytical efforts is functionally illustrated by the work of Yang et al. (Yang et al.,
2018). The authors modeled media representing polydisperse fibers that also incorporated simulated polydisperse particles in order to
develop a predictive model for the pressure drop. Furthermore, a dimensionless group was related to the moments of fiber and particle
distribution functions, and in their investigation for a predictive analytical expression found that the most effective moment was the
diameter of a monodisperse fiber distribution with an equivalent surface area and solidity as the generated fibrous geometry. This is a
reasonable claim, as the pressure drop through the media will be dependent upon drag generated by the fibers and thus an expression
utilizing the surface area and solidity is a logical conclusion. This, study not only provides insight into the use of equivalent quantities
in analytical expressions, but also illustrates the advantage of applying computational models in conjunction with other methods.
Furthermore, this model also illustrates a topic that has been addressed earlier, being the increased flow and particle penetration
through channels formed between fibers. This concept is visually illustrated below in Fig. 13.

4.3. Computational models for Particle cake formation

Investigation of particle capture mechanisms and subsequent filter loading, such as the evolution from depth loading to surface
loading, has presented a modeling challenge. Many analytical models, such as the Bergman model noted in the previous sections,
characterize particle loading into three regimes but do not explicitly model how the regimes function together. Other models, such as
the Novick model discussed in later sections, begin their analysis at surface loading by assuming negligible depth loading. The
Payatakes model noted above sought to explain how particles begin loading onto fibers and the creation of particle dendrites but
stopped short of offering an analytical model that described an entire loading curve. Computational fluid dynamics allows for the
simulation of an entire loading curve, from clean filter media to fully loaded media at a filter’s end of service life.

Fig. 13. Particle trajectories and capture through 2-D random geometry. (a) Without particle diffusion, (b) with particle diffusion (Hosseini &
Tafreshi, 2010).

21
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

Deshpande et al. (Deshpande et al., 2020) investigated the growth of the particle cake during the filtration of suspensions with
non-spherical particles. This was accomplished using a multi-sphere model to simulate the non-spherical particles. It should be noted
here that this multi-sphere approach may not be representative of many particle types, such as soot particles, as the model simply
utilizes spheres overlapping to various degrees to simulate different particle shapes. However, especially in a lab setting where test
particles are controlled, this work may estimate performance well if the sphericity of the particle is high. The void fraction of the
particle cake increased with increasing sphericity, while the void fraction decreased with a decrease in the coefficient of sliding
friction. This was attributed to particles sliding and rolling over each other, forming a denser cake rather than forming bridge-like
structures. A larger pressure drop was also attributed to non-sphericity in particle beds with nearly equal void fractions, and it was
noted that the cake consolidated when the Reynolds number was sufficiently high. Finally, the study reported that the pressure drop
increased with an increase in the unit length of the particles and decrease in sphericity to a point, but subsequently decreased as the
sphericity decreased past that point. This was attributed to an increase in the void fraction with a decrease in sphericity, which affected
the pressure drop more than the effect from the decrease in sphericity. This work offers an insight into an application of CFD that may
be applied to the study of particle cake formation and performance, such as has been discussed in the above section regarding the fiber
geometry. Potentially, this may be used to address traditionally difficult issues, such as the value of the porosity for the particle cake
from different particle types and distributions. Shu et al. (Shu et al., 2020) discuss the use of OpenFOAM as a cost-effective and
open-source CFD software to investigate gas-solid flow characteristics in fibrous media. The authors simulated particle deposition and
the formation particle dendrites immediately prior to the formation of the particle cake using a Eulerian-Lagrangian method. Their
results were favorably comparable to results generated using a CFD Discrete Element Method (CFD-DEM) in ANSYS Fluent.
Furthermore, the method used by the authors removes the need to search for adjacent particles, enabling a reduction in computational
time. This was illustrated by the authors, who stated that the OpenFOAM simulations required about half as much time as the sim­
ulations completed using Fluent. This touches upon an important aspect of work utilizing CFD simulations, namely the amount of time
required and the potential impact that this has on the research work itself. A fast yet accurate model and methodology will assist in the
future application of CFD simulations and will make them more feasible for use with relatively modest computer hardware and
therefore accessible on a larger scale as a research tool.

4.4. Voxel-based computational fluid dynamics models

Computational modelling makes filtration simulations and predictions more accessible on a day-by-day basis. However due to the
complex nature of aerosol filtration, simulations can be prohibitively time consuming and computationally expensive, as mentioned in
the previous section. Scientists and engineers continue to render computational models that are sufficiently accurate while also
improving the speed and cost. An example of such an advancement is the GeoDict software. From 2001 through 2011, a team from
Fraunhofer Institute for Industrial Mathematics (ITWM) developed GeoDictfor microstructure simulations. (Math2Market, n.d.).
GeoDict is an extensive digital material laboratory for composites, fibers, foams, ceramics, and metals (GeoDict, 2019), and utilizes a
voxel-based model instead of a mesh-based model, offering multi-scale modeling down to nanometer range, visualization of material
structures, and computational simulations of material behaviors under a range of conditions. In the development of digital twins for
the modelling of fibrous filter media, this software provides multiple tools to produce high-fidelity replicas. Real filter media can be
digitally replicated in GeoDict by importing SEM images or CT scans of the media using the ImportGeo-Vol module of FiberGeo, and
applying image filtering techniques (Linden et al., 2018). First, ImportGeo-Vol module extracts statistical characteristics from the filter
media such as fiber size distribution, fiber orientation, filter solidity, etc. Second, the FiberGeo module creates digital twin replicas of
filter media from the statistical characteristics, whether generated from the scanned images or provided directly by the user. GeoDict
software enables computational analysis of filter permeability and flow resistance through the FlowDict module (Azimian et al., 2020).
Finally, the FilterDict-Media module of GeoDict simulates the transport, collision, and deposition of aerosol particles through the
digital twin media over time. Overall the GeoDict software package offers a comprehensive solution for computational modeling of air
filter media, from the production of digital twin replicas to the simulation and analysis of the particle capture, filter efficiency, pressure
drop, and particulate cake and deposit formation through varied operating conditions (Alilou, 2018; Bai et al., 2020; Maddineni et al.,
2018).
Although GeoDict offers an alternative to more traditional CFD options such as Fluent, several question may be raised on the
application of a voxel-based model. Hoch et al. (Hoch et al., 2020) created and validated mesh-based and voxel-based models using
liquid droplets for comparison against each other, collected experimental data, and the single-fiber model. GeoDict was selected for the
voxel-based model, whereas StarCCM+ was selected for the mesh-based model. The comparison of the single-fiber model against a
single digital fiber showed the MPPS was similar from the voxel-based model, although the resulting efficiencies were different by
several percentage points over the entire size range. The mesh-based model yielded similar efficiencies over a range of very small
submicron particles but deviated and was consistent with that of the voxel model for a majority of the particle size range investigated.
For a comparison against experimental data, a representative portion of the test media was analyzed using nano-CT scans and imported
into a digital model. The results of the simulations were extrapolated to estimate the performance a whole filter. For 0.25 μm particles,
the capture efficiencies were close for both CFD models, and the pressure drop was similar but varied within a range of around 200Pa,
where the mesh-based model yielded closer results. The authors concluded that the deviation from the pressure drop results was likely
due to the extrapolation of the simulation results, which would implicitly assume that the entirety of the filter is comprised of media
with the same characteristics. This is unlikely, making the explanation reasonable, as a similar phenomenon has also been discussed in
above sections regarding media inhomogeneity and its effect on performance. Furthermore, since the size range of particles used in the
study is 1 μm and less, the dominating capture mechanisms will be diffusion and interception. This offers a simple explanation for the

22
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

similarity from the CFD results to the experimental results regarding the capture efficiency, specifically regarding the impact of media
nonuniformity on diffusion. Although this study considers liquid droplets, the results should provide a reasonable estimation for how
the mesh-based and voxel-based models will compare to each other using solid particles. As a final note, it must be mentioned that the
computational time reported for the mesh-based model was exceedingly long, and orders of magnitude greater than the necessary time
for the voxel-based approach, both using two 2.6 GHz processors. This clearly highlights that a benefit of using the voxel-based model
is the significant decrease in computational time for a model of comparable results to a mesh-based model, expediting the research
process and potentially allowing for a greater scope of work.
Due to the nature of the voxel-based model, a distinction should be made between resolved and unresolved models. Resolved
models are defined as a grid that has a resolution capable of resolving the fibers, i.e., fibers and particles are comprised of a least one
voxel that is considered to be solid, implying that the voxels are smaller than the smallest fiber or particle. On the other hand, an
unresolved model implies that the voxels are larger than the smallest fiber or particle, thus leading to the conclusion that an alternative
method of tracking the fibers and particles is necessary. This inevitably leads to the question of how a voxel is counted as solid. For the
resolved models, a typical choice would be to consider the voxel as solid if half of its volume is occupied. Thus, all voxels are either
considered to be solid or empty. However, for an unresolved model, a voxel may be considered as a porous medium if a particle or fiber
is not resolved, where the Navier-Stokes-Brinkman equation is solved to yield the solution. Becker et al. (Becker et al., 2016) inves­
tigated this multiscale issue between resolved and unresolved models for cake formation by simulating a resolved particle loading
model, an unresolved particle loading model, and models at intermediate stages in between. A fully resolved model was used to
determine the maximum solidity and resistivity of the resulting particle cake. These characteristics were used as inputs for the fully
unresolved model, namely the local maximums for each voxel, and showed good agreement with the resolved model. However, the
models at intermediate levels of resolution required adjusted inputs which followed linearly increasing relationships as model reso­
lutions approached fully resolved from the initial inputs derived from the overall particle cake characteristics. This is consistent with
the expectation, as the solidity and resistivity of a solid voxel are 1 and infinity, respectively, meaning that the smaller the voxel
resolution was, the closer to these values that local maximums approached to achieve similar results. This work effectively illustrates
the validity of multiscale modeling, although care must be taken to ensure that the proper input parameters are used to achieve good

Table 3
Notable experimental air filtration models.
Author Details

Bourrous et al. (Bourrous, Bouilloux, Nerisson, Thomas, & Appert-Collin, 2017) Effects of pleat geometry, relationship of deformation and
air flow resistance
Brown and Thorpe (Brown & Thorpe, 2001) Effects of fiber diameter size distributions on efficiency
and flow resistance
Humphries (Humphries, 1980) Effects of packing density uniformity in woven fabric
filters
Dhaniyala and Liu (Dhaniyala & Liu, 2001a) Effects of homogeneity of filter media on filtration
efficiency and resistance
Leung and Hung (Leung & Hung, 2008) Effects of fiber diameters, comparison of microfiber and
nanofiber filtration
Penicot et al. (Penicot, Thomas, Contal, Leclerc, & Vendel, 1999) Effects of solid and liquid particle loading on HEPA filber
media
Artous et al. (Artous, Bouilloux, Marchal, & Ouf, 2012) Effects of clogging process in mini-pleated filters,
applicability of Darcy’s law
Godoy and Thomas (Godoy & Thomas, 2020) Effects of relative humidity on clean air filter media
Novick et al. (Novick, Higgins, Dierkschiede, Abrahamson, & Richardson, 1991; Novick, Klassen, Effects of particle depth loading, variation of Bergman
et al., 1992; Novick, Monson, & Ellison, 1992) model
Song et al. (Song, Lee, Park, & Lee, 2007) Effects of particle size distributions on pressure drop
Aguiar and Coury (Aguiar & Coury, 1996) Effects of surface loading of particle cake of limestone
particles
Cheng and Tsia (Cheng & Tsai, 1998) Experimentally determined particle cake thickness and
porosity
Chen and Hsiau (Chen & Hsiau, 2009) Measured and analyzed the distribution of particle cake
thickness
Koch and Krammer (Koch & Krammer, 2008) Analyzed inhomogeneous media permeability and
resulting flow velocities
Endo et al. (Endo, Chen, & Pui, 1998b) Measured particle cake thickness in situ, and addressed
cake compression
Endo et al. (Endo, Chen, & Pui, 1998a) Effects of a bimodal aerosol distribution on the particle
cake porosity
Kim et al. (Kim, Wang, Shin, Scheckman, & Pui, 2009) Loading of soot particles, accounting for particle shape
and size polydispersity
Sun et al. (Sun et al., 2019) Effects of particle shapes, NaCl particles and soot
agglomerates
Bourrous et al. (Bourrous et al., 2016) Particle clogging of media, considering effects media
characteristics
Joubert et al. (Joubert et al., 2011) Investigated the influence of humidity on surface
filtration

23
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

results. Overall, the results of this study suggest that a coarser, unresolved grid may yield comparable results to finer, resolved grids.
This will inevitably decrease the necessary computational time of a model and offers more options for future work, specifically
considering particle size distributions for nanoparticles. However, this must be tempered with the fact that this model represents a
hybrid approach since the unresolved grid utilized inputs from the resolved grid, suggesting that further refinement is necessary for a
standalone application of an unresolved model.
Regarding an extension of multiscale modeling, Lee et al. (Lee et al., 2021) explored a sequential approach to analyze the per­
formance of a facepiece using Xμ-CT scans to import the media geometry into a digital model. The term sequential was used to describe
this work, as the media performance was estimated on the microscale by simulating a representative portion of media and extrapo­
lating the performance of the entire facepiece using this information. A comparison between the model and experimental results
showed good agreement for the representative media portion, although the penetration was stated as lower for the experiments until
the full mass loading was achieved. The airflow resistance agreed well over the entire range of mass loading. Regarding the full
facepiece, the simulation and the experiment agreed well, implying that this methodology reasonably approximates macroscale
performance from an analysis of a microscale model. This is similar to the aforementioned point from Saleh et al. (Saleh et al., 2014)
regarding the calibration of a macroscale model using a microscale model. However, it should be noted that this work is somewhat in
contradiction to previously mentioned work regarding the extrapolation of a representative portion of media to the full-scale, due to
media inhomogeneity. This work suggests that the method of sequential multiscale modeling may effectively be applied to complex
situations, such as the filtration of a facepiece or pleated filter with reasonable accuracy.

5. Experimental modeling of air filtration

Experimental work in the field of air filtration covers a broad range of topics. These methods offer a macroscale approach to
answering specific questions that are difficult or impossible to answer from analytical or CFD modeling. Experimental modeling of air
filtration simplifies the determination of specific parameters that are too complex to accurately determine analytically, allowing for the
empirical determination of necessary constants and correlations. Significant efforts in experimental modeling of air filtration are
shown in Table 3.

5.1. Effects of pleated geometry and air flow resistance

Bourrous et al. (Bourrous et al., 2017) investigated the relationship between pleat deformation and air flow resistance. The pleat
deformation was induced by increasing the clean air flow rate through the filter, showing that the pressure drop over the induced air
flow range did not follow Darcy’s law. Higher flow rates created contact between the inner surfaces of the pleats on the downstream
side which led to a reduction of the air flow area. To address this result, the authors introduced a correction factor for the observed
nonlinear behavior. It is interesting to note that for the geometry considered in the study, the pressure drop from very low flow ve­
locities below 0.05 m/s did not deviate from Darcy’s law. However larger flow velocities above 0.03 m/s created a notable loss in the
surface area. The deviation from Darcy’s law is a notable result, as a significant portion of modeling work uses this law either directly
or indirectly. Pleated models that don’t incorporate or acknowledge this may suffer from inaccurate results, although empirical work
may implicitly account for this via correction factors. In other work, Bourrous et al. (Bourrous et al., 2014) experimentally modeled the
loading of flat-sheet and pleated HEPA media with nanoparticles to analyze the influence of the pleat geometry on the distribution of
particles within the media. It was noted that for the depth loading of the particles, no functional difference was found between
flat-sheet media and pleated media. This infers that for filter depth loading using nanoparticles, flat-sheet studies and models could be
applied to pleated media. While the analytical work of Fotovati et al. (Fotovati et al., 2011) critiqued the application of semi-analytical
expressions developed from flat-sheet media to pleated media, the experimental modeling of Bourrous suggested this is indeed possible
to some degree.

5.2. Effects of fiber diameter size distributions

Experimental modeling has also been useful in investigating the geometry and structure of the fibrous media. Brown and Thorpe
(Brown & Thorpe, 2001) conducted experiments to investigate the influence of a bimodal fiber size distribution within the media, and
the effects on the performance of the filtration media using different mixtures of fiber sizes. It is noteworthy that the authors inves­
tigated whether the number-weighted, surface area-weighted, or volume-weighted mean most accurately described the behavior of the
fiber mixtures. This is relevant to the earlier discussion in previous sections regarding the efficacy of representative and mean diameter
values for predictive use in analytical models. The results of the experiments indicated that the surface area-weighted mean best
described the observed data, which is a similar conclusions drawn from aforementioned work (Yang et al., 2018). Furthermore, the
particle penetration was noted to decrease with a decreasing mean fiber diameter size, which was expected.

5.3. Uniformity of packing density

Humphries (Humphries, 1980) experimentally modeled the relationship between air velocity and air flow resistance to the mean
and variance of the fiber packing density of clean filter media. However, it is worth noting that this work focused on woven fabric
filters instead of nonwoven fibrous filters. The author stated that the mean packing density was not sufficient to predict the particle
collection efficiency of an air filter, where a wider distribution in the packing density resulted in lower air flow resistance. Thus, it was

24
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

suggested that a filter having a higher degree of non-uniformity would result in a lower flow resistance due to the non-uniform velocity
distribution. A noteworthy conclusion from this study is that although the areas of the filter media experiencing high flow velocities
were relatively small compared to the areas experiencing lower flow velocities, the volumetric flow through the high-velocity areas
was substantially larger and represented a substantial fraction of the total air flow. From this, it may be inferred that there is an
increased chance of particle capture in areas experiencing high flow velocities, and that different capture mechanisms may be
dominant in different local areas of the filter media. This is consistent with previously discussed analytical and computational models.
Dhaniyala and Liu (Dhaniyala & Liu, 2001a) investigated the capture efficiency of filter media related to local packing densities
created by particle loading and subsequently modeled the inhomogeneity. It was interesting to note that the authors paid great
attention to creating a flat velocity profile through the filter media instead of the normal parabolic flow profile in order to prevent a
non-uniform particle count at different locales. The results of their experiment indicated that the upstream concentrations followed a
Poisson distribution, while the downstream concentration followed a Gaussian distribution, thus indicating media inhomogeneity
since the distribution changed. To address this, a non-uniformity factor was proposed to account for the inhomogeneity that would
function to decrease the theoretical single-fiber efficiency. It is interesting to note that the non-uniformity of the downstream dis­
tribution decreased with an increasing particle size, but that the variance in the data was prohibitive in allowing for a relationship to be
defined. Furthermore, it was noted that the non-uniformity factor was a function of the face velocity, where an increase in the face
velocity resulted in a narrower distribution of the local capture efficiencies. Low-efficiency filters were shown to result in lower
non-uniformity factors, while high-efficiency filters had larger non-uniformity factors under the subjected experimental conditions.
Finally, the authors suggested that a decrease in the Peclet number resulted in a decrease in the non-uniformity factor, and the
non-uniformity factor seemed to be independent of the interception parameter. An implication that may be inferred from this work is
that nonuniformity of the filtration media may be estimated based upon variables that are relatively easy to measure, such as the
particle size distributions upstream and downstream of the media.

5.4. Effects of microfiber and nanofiber media

Leung and Hung (Leung & Hung, 2008) conducted experimental investigations on the filtration efficiency of microfiber and
nanofiber media operating in the aerodynamic slip regime. The outcome and findings of their experimental results, shown in Fig. 14,
resulted in new semi-empirical models of air filtration efficiency and flow resistance. The authors noted that when particles deposit
onto fibers and form dendrites, new filter geometry including both fibers and particle dendrites may be considered as a binary mixture
of two different fiber types. This outcome aligns with the work of Brown and Thorpe (Brown & Thorpe, 2001) and is similar to the
previously discussed analytical models. Leung and Hung noted that the nanofiber media investigated were more capable of particle
capture than the microfiber media, which was attributed to the increase in particle deposition by interception. Furthermore, the
nanofiber media were shown to have a smaller shift in the MPPS than the microfiber media, which was attributed to the fact that the
nanofiber media was already comprised of smaller fibers, and already favored deposition by interception from the large
surface-to-volume ratio as compared to the microfiber media. However, the nanofiber filters were noted to be more efficient but not
necessarily better than microfiber filters since the microfiber filters were capable of higher particle mass loading than the nanofiber
filters. It may be inferred from this work that a decision may be made based upon either application or necessity to cater the char­
acteristics of the filter media to a specific situation. For example, a microfiber media may be better suited for a low-efficiency

Fig. 14. Filter efficiency evolution of filter samples taken at single hour intervals (Leung & Hung, 2008).

25
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

application where it is more desirable to extend the lifetime of the filter, i.e., more mass loading, than it would be to ensure smaller
particles are captured.

5.5. Effects of solid and liquid particle loading

Penicot et al. (Penicot et al., 1999) studied the effects of solid and liquid particle loading on HEPA filter media. Electron micro­
graphs of the loaded filter media showed that the loading could be segregated into depth loading and surface loading, consistent with
commonly accepted air filtration theories. From this work, the authors proposed a relationship between the particle diameter dp and
the transition point from depth to surface loading wT (with units of g/m2) as shown by Equation (28).
wT = 3.975 · dp (28)

The mass-loading point proposed by wT is intended to be the intersection of the tangents between the two linear portions of the
curve defined by the depth and surface loading. It is interesting to note that the filtration velocity did not affect the transition point,
although this was not investigated at different particle sizes. The porosity of the particle cake was estimated using three different
methods: (1) loading a significant amount of particles onto the filter media then measuring the thickness of the cake, (2) calculating the
slope of the pressure drop curve then back-calculating the porosity of the cake using the correlation proposed by Davies (Davies, 1953),
and (3) simulating the filtration process and the formation of the dendrites. The authors noted good agreement with all three methods
and previous studies. Thus, experimental modeling proved valuable for estimating historically difficult yet functionally important
variables for particle cake porosity. The results and methods presented in this work may easily be incorporated into other analytical
models, which typically focus on a single loading regime. As briefly mentioned in section 2, the transition point may theoretically be
used to distinguish where a model for one loading regime ends, and another beings, effectively enabling the investigation of the entire
loading process (Thomas et al., 2001).

5.6. Filter pleat clogging and particle cake porosity

Artous et al. (Artous et al., 2012) investigated the clogging of mini-pleat filters and resulting particle cake porosity using graphite
nanoparticle aggregates. It is noteworthy that the authors used the Novick-Kozeny model, discussed below, and optical measurements.
The results of the experiment indicated that the clogging process was independent of the filtration velocity for flat filters, which is
consistent with the work of Joubert (Joubert et al., 2011). Below the saturation point, humidity was not an influencing factor,
however, particles became condensation nuclei and essentially liquid aerosols when the humidity was above the saturation point. It is
noteworthy that the authors used a value of twice the clean filter pressure drop to indicate when the loading regime changed from
depth loading to surface loading, illustrating a simple method of estimating the change of loading regime. Furthermore, the authors
noted a nonlinear evolution in the pressure drop at higher velocities and hypothesized that Darcy’s law could not describe the flow past
2.9 cm/s. This potentially presents issue with many analytical models due to the prevalence of Darcy’s law in their derivations, as
mentioned above in section 5.1 with the work of Bourrous et al. (Bourrous et al., 2014). The authors noted that the pressure drop was
highly sensitive to the porosity, which may be influenced by different variable estimations. Finally, the authors observed for velocities
lower than the analyzed velocities in their experiments that flat filter geometry may eventually adequately represent mini-pleat ge­
ometry during the surface filtration stage of filter clogging for varied air flow geometries. This is in contrast to the statements addressed
earlier (Fotovati et al., 2011) regarding the inability of flat sheet models to describe pleated media.
On the subject of relative humidity, Godoy and Thomas (Godoy & Thomas, 2020) investigated its influence on the loading of soot
particle onto flat-sheet media. Specifically, the media was loaded under dry air conditions to a defined pressure and permeation.
Afterwards, the media was subjected to moist air and analyzed for changes to pressure and permeation before being removed and
dried. The testing-drying process was repeated three times, and it was experimentally shown that the pressure drop was not perma­
nently affected by changes in humidity. Although a seemingly straightforward conclusion, it is suggested that the importance of this
work lies in the realistic operation of air filters, where the environment may not always be controlled outside of a lab setting. This
implies that exposure to an environment with varying degrees of humidity may be inevitable, where a temporary change of perfor­
mance may be expected.

5.7. Determination of resistance factors

Novick et al. (Novick et al., 1991) proposed an alternative approach to the Bergman model, where the total pressure drop ΔP is
considered to be the sum of the pressure drop from the clean filter ΔPo and the pressure drop across the particle cake ΔPp as shown in
Equation (29). In this equation, K1 and K2 represent specific resistance constants dependent on the physical characteristics of the filter
media and particle cake, respectively, v is the gas velocity, m is the particle mass loading, and A is the active filtration area.
m
ΔP = ΔPo + ΔPp = K1 v + K2 v (29)
A
The constant K2 may be developed theoretically from Stoke’s Law, however this method assumes that each particle is not influ­
enced by its neighboring particles. As an alternative, an expression for K2 may combine a resistance factor RN to account for the affect
that packing density of the particles on the overall pressure drop. Several theoretical and analytical methods attempt to determine the
resistance factor, however it may be experimentally determined from the given equations as shown in Equation (30).

26
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

ACc d2 ρp (ΔP − ΔPo )


RN = (30)
18μvm
From experiments conducted using NaCl particles with varying particle size distributions, the authors illustrated a strong
dependence of the resistance factor on the particle diameter, as well as the linear nature of increased pressure drop as particle mass is
loaded onto the media. Curve-fitting the data gathered from their experiments yielded an empirical expression for the resistance factor,
giving the pressure drop across the particle cake due to the deposited mass as shown in Equation (31).
( )
mμv 3.06 + 7.56 · 107 Cc d2 ρp
ΔP − ΔPo = 18 (31)
ACc d2 ρp

However, it should be noted that this was developed specifically for solid and dry particles using the centimeter-gram-second
system of units, where the cited work stated that it should be applicable for particle shapes with an aspect ratio close to unity.
Finally, Equation (31) may not be applicable to particles that are sticky or have high adhesive coefficients such as very small particles,
as this would cause the particle cake to form differently and cause variations in the porosity. This work effectively illustrates an
empirical method to estimate the porosity of the of the particle cake, which may be coupled with CFD methods as discussed earlier for
calibration or other investigative purposes. However, it should be noted that the use of this correlation and model may oversimplify the
dynamic properties of the particle cake, which is discussed further in section 5.9.
Similar to the work of Novick et al., Song et al. (Song et al., 2007) studied the evolution of the pressure drop in fibrous HEPA filters
as a function of loaded mass with respect to the aerodynamic particle size. Of particular interest is the experimental focus on a
polydisperse aerosol size distribution. Furthermore, it was stated by the authors that the filters that were manufactured with an acrylic
binder yielded more scatter in the particle collection data than filters that were manufactured without acrylic binder. This was
attributed to the increase in fiber thickness caused by application of the acrylic binder, thus creating a different deposition pattern
favoring particle loading into the depth of the media. The authors noted the porosity of the particle cake was only slightly influenced by
the characteristics of the filter media itself, although the porosity of the cake was strongly correlated to the particle diameter
comprising the cake. New empirical equations were presented as the result of the experimental work of the authors, and predicted the
pressure drop evolution as a function of the aerodynamic particle diameter, da, of a monodisperse particle distribution and the geo­
metric mean diameter of a polydisperse distribution, dg, given below in Equations (32) and (33), respectively. It is interesting to note
the similarity of Equation (32) to the model proposed by Novick in Equation (35). However, the use of the geometric standard de­
viation to address particle polydispersity suggests that Equation (33) may be a better candidate for test particle and realistic appli­
cations which are not subjected to a monodisperse particle distribution. Given the nature of the correlations, it is important that the
prescribed units are utilized with velocity expressed in meters per second, mass expressed in kg, diameters expressed in meters,
resulting in a pressure drop expressed in Pascals.
(( ) )
1.248 vm
ΔP − ΔP0 = − 1.22 · 105 (32)
da A
(( ) ( )
1.248 ( )2 ) vm
ΔP − ΔP0 = exp − 0.5ln σ g − 1.22 · 105 (33)
dg A

5.8. Particle loading in pleated filters

Another alternative correlation proposed by Novick et al. (Novick et al., 1992) uses the model proposed by Bergman et al. as the
foundation to account for particle loadings within the filter depth. Beginning from the Bergman model, and using the aforementioned
method from the same author above, K2 can be determined as shown in Equation (34):
√̅̅̅̅̅
αf
K2 = 64μ (34)
df dpf ρpd

However, this expression still requires the knowledge of media and particle characteristics that are difficult to obtain. Therefore,
experiments were conducted to correlate K2 to the particle MMD, where data were curve-fit in order to give a reasonable prediction for
K2 as a function of particle mass median diameter shown in Equation (35). Given the correlated nature of Equation (35), it is worth
noting that the airflow resistance is given in units of Pa, area in m2, velocity in m/s, MMD in meters, and mass in kg.
ΔP − ΔPo 0.963
m = − 1.64 · 105 (35)
A
v MMD

This correlation is only valid for dry and solid particles, and application for pleated filters should be limited to cake thickness less
than half of the spacing between the layers of filter material to avoid any bridging that may occur between the pleats if enough mass is
loaded onto the filter. Particle cake bridging between layers would begin the reduction in filtration area, as discussed in previous
sections. These correlations were created to circumvent the determination of problematic parameters, such as the particle cake
porosity and the effects of the particle diameter distribution and serves as a practical method of performance prediction. The Novick
models have been commonly used in other analytical work that focuses on solid particles in order to link the different loading regimes
and provide an estimate for the full spectrum of particle loading (Thomas et al., 2001).

27
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

5.9. Surface loading and properties of the particle cake

Many experimental investigations have studied the effects of particle deposition in the depth of filtration media. However, surface
loading is equally as vital to the full understanding of the particle loading process and for some applications constitutes the vast
majority of a filter’s life. Aguiar and Coury (Aguiar & Coury, 1996) investigated the porosity created by loading limestone particles
onto flat-sheet fabric filters. The authors found that both the Ergun equation as well as the alternative Rudnick-Happel equation, shown
below in Equation (36), yielded reasonable agreement to each other, although neither equation showed good agreement at a cake
height near the cake-surface interface of the media.
( 5
)
18Qμvt 3 + 2(1 − ε)3
ΔP = (36)
Aρp dp2 3 − 4.5(1 − ε)13 + 4.5(1 − ε)53 − 2(1 − ε)2

The experiments of Aguiar and Coury showed the porosity tended to decrease as the cake height increased, and also decreased over
time which indicated compression of the particle cake. This constitutes a potential issue for the analysis of particle cake performance,
as many models consider the porosity of the particle cake to be static, similar to the porosity of the fibrous media without regarding the
effects of particles loading into the media depth. To address this, the authors proposed a new form of the Ergun equation. Cheng and
Tsai (Cheng & Tsai, 1998) experimentally evaluated the specific resistance of the particle cake using a laser displacement system to
determine the cake thickness and porosity and compare them to theoretical predictions. It was stated that the airflow resistance may be
considered as the sum of the resistance from the particle cake and clean filter medium, and that the specific cake resistance increased as
the particle size decreased, as is expected. Furthermore, for all the particles tested, a higher filtration velocity resulted in a larger
resistance and lower cake porosity. This is a notable result, as it suggests that the particle cake may have different physical charac­
teristics and thus perform differently for the same kind of particles based upon the filtration conditions, not to be confused with a
change in particle collection mechanisms of the cake at a given point in time. Particle shape was also attributed to have an impact on
the porosity of the cake, where it was mentioned that more irregular particles could possibly lead to more random arrangements inside
of the cake, resulting in a more porous structure. The compressibility of the particle cake was also investigated, where it is interesting to
note that the cake was considered somewhat elastic, and that the loading structure expanded with a decrease in filtration velocity at a
constant mass loading. As filtration velocities were decreased from higher loading velocities to a velocity of 1 cm/s, the slope of the
airflow resistance was greater than when the loading velocity was decreased from a lower magnitude to 1 cm/s. This was attributed to
the higher velocities resulting in a more compact particle cake, and also the elastic expansion when the velocity was decreased. The
authors noted that spherical particles seemed to experience greater cake compressions than particles of more irregular shapes in the
tested ranges of filtration velocity. Chen and Hsiau (Chen & Hsiau, 2009) introduced a multipoint measurement method to measure the
average thickness of the particle cake, resulting in a suggested non-uniform thickness distribution. Furthermore, the authors also
investigated the impact of the particle cake and its characteristics on the performance of the filter overall, specifically noting the
impact that different filtration conditions had on the formation of the cake, and its performance. The experiments were performed
using dust taken from a coal-fired power plant, with a gaussian distribution between 0.55 μm and 18.17 μm, a mass median aero­
dynamic diameter of 8.31 μm, a standard deviation of 2.04. As in the previously cited work, the Ergun equation was used to estimate
the cake porosity. To measure the thickness of the particle cake, a pressure sensor was used to determine when contact was made. This
method was stated to have an accuracy of 98.26% with a standard deviation of 0.050 mm when measuring particles comprised of
diameters smaller than 1.5 mm. The experimental results indicated that higher filtration velocities resulted in larger pressure drops
from the same particles, and that the maximum airflow resistance occurs when the porosity is at its lowest, which are both expected
results. Regarding particle cake compression, the authors discussed that the compressive stress acting on dust particles could be
separated into normal and shear components, and that the dust cake would compress and decrease in thickness when the effect of the
shear component becomes larger than the normal component. It was noted that when the cake compressed, that the airflow resistance
dropped, as well as the collection efficiency. The drop in airflow resistance suggested that, for a time, the airflow resistance could
decrease before more particles were loaded, resulting in a further increase to cake thickness and airflow resistance. The concept of
particle cake compression is somewhat ignored in many models, where it is assumed that the porosity is static and a function of an
equivalent particle diameter. However, it may be implicitly incorporated, especially for empirical models, if the effect is somewhat
negligible, for example with a specific kind of particle or at relatively low filtration velocities. Somewhat related to the previously cited
work is the work of Koch and Krammer (Koch & Krammer, 2008), who proposed a model based upon Darcy’s law to describe an
inhomogeneous filter medium permeability. Of particular interest was that the authors suggested that areas of the filter with a lower
permeability would also have a lower filtration velocity, and thus would experience a slower formation of the particle cake as opposed
to areas of the filter medium with a higher permeability. This corroborates previously discussed work related to the permeability and
flow velocity, although it is interesting to note that the formation of the particle cake may be affected resulting in different local
porosities or even cake compression.
Endo et al. (Endo et al., 1998b) studied the dust loading process and performed in situ measurements of particle cake thickness. The
authors noted a dependence of the airflow resistance on velocity as is expected, however the height of the cake was noted to be almost
independent of the filtration velocity. Furthermore, the compression of the particle cake was stated to occur under very similar
conditions for all of the different particles tested. It was stated that compression of the particle cake would occur when the drag force
acting on a specific layer of particles becomes larger than the strength of the cake layer. Specifically, it was stated that the strength of
each layer was inversely proportional to the diameter of the particles comprising that layer, and therefore layers of smaller particles are
less easily compressed than layers of larger particles. The authors stated that the Kozeny-Carman equation was only applicable to

28
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

monodisperse and fixed-bed situations, which may be acceptable for experimental work, but may not be applicable for realistic ap­
plications. Endo et al. (Endo et al., 1998a) continued their work and considered the loading of an aerosol with a bimodal distribution.
Of particular interest is an analytical expression given in their proposed model that may be used to calculate the porosity of the particle
cake for a bimodal distribution given the known concentration of particles n1 and n2, shown below in Equation (37).
( ) ( )
π ( ) 9 ( ) π ( ) 9 ( )
α = n1 3 ln dvg1 + ln 2 σ g1 + n2 3 ln dvg2 + ln 2 σg2 (37)
6 2 6 2

It is important to note that the equation is still valid is the distribution is unimodal, and n2 is equal to zero. The authors experimented
with mixtures of Al(OH)3 and Arizona Road Dust (ARD), and found that the airflow resistance for the mixture of particles was between
the resistance for cases with just one kind of particle or the other. Specifically, Al(OH)3 are smaller particles than ARD, and therefore
the airflow resistance increased as the ratio of Al(OH)3 to ARD increased, as may be expected when considering the airflow resistance
of smaller particles at equivalent mass loadings to larger particles. However, when analyzing the particle cake height, it was shown that
a case with 10% ARD mixture had a significant reduction in height compared to a case with a 50% mixture. A possible explanation was
given, suggesting that the smaller Al(OH)3 particles may have filled the voids between the larger ARD particles instead of contributing
to the increase in cake height, resulting in a cake with a lower porosity. The application of this model is complicated by the necessity of
estimating the number of particles present within the particle cake, limiting its practical use. However, it may be possible for a
reasonable estimate based on the use of a transition point between the depth loading and surface loading. If the transition point is
known, then the mass loaded into the depth may be separated from the subsequently loaded mass, enabling an estimation of the
particles comprising the cake.
Kim et al. (S. C. Kim et al., 2009) characterized the loading of soot agglomerates onto a fibrous filter, using the work proposed by
Endo et al. to model the loading using the primary particle size distribution of the particles. The authors noted that the agglomerated
particles have a lower penetration compared to other particle shapes, such as spherical particles, with the same mobility diameter due
to their shape and large interception length. The resulting particle cake porosity was noted as relatively high compared to other cake
porosities from literature, where this was attributed to the loose structure from the agglomerates given that the resulting particle cake
formed was not easily compressible. It was stated that the pressure drop model presented by Endo et al. accounting for the particle
shape and polydispersity of the aerosol was appropriate for the modeling of the mass loaded onto the filter. Sun et al. (Sun et al., 2019)
investigated filter performance and loading characteristics using two kinds of composite filters with different structures by challenging
them with NaCl particles and soot agglomerates. The authors measured the particulate mass collected by the filter by incrementally
weighing the filters during the loading experiments, where the final weight was determined for a corresponding pressure drop of five
times the initial pressure drop. It was noted by the authors that NaCl particles and soot agglomerates loaded into the filters differently
due to their particle shapes, which is a reasonable and expected result. More compact particle cakes were formed at higher filtration
velocities, resulting in higher air flow resistance from the compressed cake, corroborating the previously mentioned work regarding
cake compression.
Bourrous et al. (Bourrous et al., 2016) studied the clogging of fibrous media challenged with nanoparticles, considering the
penetration of the particles into the media and porosity of the particle cake, as well as the physical characteristics of the filter medium
and challenge aerosol. The authors used a model based on a capillary analogy instead of the typical model based upon Darcy’s law and
stated that it resulted in a close agreement in terms of the transition point and airflow resistance for their experiments with high
efficiency filters and submicron particles. However, the results of the experiment diverged somewhat during the surface loading
regime. This was attributed to the influence that the filtration media exerts on the formation of the particle cake. Thus, an interesting
point has been introduced implying that the media itself has some impact on the formation of the particle cake, although this concept
does not seem to have been investigated thoroughly in current literature. It noteworthy that authors used a combined approach to
determine the pressure drop for a partially clogged capillary by noting the flow passing through the clogged portion, and the flow
passing through the free portion of the capillary. Although, the model considered the capillaries to be uniform, whereas realistically the
capillaries are known to be winding and tortuous. As with previously cited work, the use of submicron particles may exacerbate the
prevalence of particulate loading into the depth of the media and should be considered for the application of this model to larger
particle sizes.
Finally, Joubert et al. (Joubert et al., 2011) conducted experiments using hygroscopic and non-hygroscopic particles and proposed
a semi-empirical model for the evolution of the pressure drop of a HEPA filter during surface filtration, including the effects of hu­
midity. Similar to the work discussed earlier by Thomas et al. the particle cake was separated into layers for evaluation. It was sug­
gested by the authors that humidity affects the structure of the particle cake by modifying the adhesive forces between the particles,
resulting in a restructuring of the cake. The particle deposits in the cake restructured to have a nonhomogeneous structure from the
creation of local preferential pathways due to the formation of aggregates, yielding a decrease in the specific resistance of the particle
cake. It is noteworthy that these effects were greater for smaller particle sizes. During their experiments, the authors noted that when a
higher humidity is maintained after particle clogging, the pressure drop across the particle cake decreases gradually as a function of
time until it reaches and equilibrium. This was explained as absorption of the water vapor onto the surface of the deposited particles in
the cake, which then restructures the deposit and reduces the specific surface area and creates preferential airflow pathways.
Furthermore, it was suggested that the effect was dependent on the level of humidity, and not the deposited mass of particles. The
proposed model had good agreement when micron-sized aluminum oxide particles were used, however the results of tests conducted
with submicron-sized sodium chloride particles agreed less, within a margin of 30% from the model prediction. This work illustrates a
noteworthy point, being the difference between experiments conducted in a lab versus possible realistic applications of filters, where

29
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

the condition in experiments are usually well-controlled and may not reflect expected operational conditions, such as areas with high
humidity. This would naturally lead to differences between the expected performance predicted from the models and the realistic
performance.

6. Discussion and conclusion

The loading process of particles onto fibrous filters is both complex and dynamic. Particles are primarily captured by the fibers
through five deposition mechanisms, which depend on the characteristics of the particles and fibers, such as diameter, density, shape,
velocity, etc. Furthermore, it is possible for a particle to bounce instead of adhering to a fiber, resulting in different loading structures
or a failure to capture the particle. As more particles are loaded onto filters, the deposited particles will form structures with the shapes
depending on the characteristics of the particles, such as their mobility, and filtration characteristics, such as the filtration velocity, or
the fiber diameter. Overall, the loading process of flat sheet media may be separated into the three loading regimes of depth loading,
transition loading, and surface loading. Beginning from clean filter media, the particles will initially load into the depth of the media,
with a linear increase in the airflow resistance. As more particles deposit onto the fibers, the probability of particle capture by pre­
viously deposited particles increases until the deposition is primarily between the particles themselves. This results in a nonlinear
increase in airflow resistance as the loading regime transitions from the depth of the media to the surface. Once the transition is
complete, particles primarily deposit to the particle cake on the surface of the media, resulting in another linear increase in airflow
resistance as more particles become captured.
Accurately modelling the loading process has been a significant and continuing effort in the research community for many decades,
resulting in many different analytical, computational, and experimental models. This work has endeavored to present and review many
of these models to provide an overarching perspective of the current state of fibrous filtration modelling. In short summary of the
different categories, several comments may be made:

• Analytical models benefit from their inherent ease of use, but rely upon assumptions and empirical factors that potentially add
implicit error, such as ordered fiber geometry, or limit their application. Besides the utilization of Darcy’s law, Bergman’s model
appears to be the most prevalent and applicable, with the latest iteration being encapsulated in Thomas’ model.
• Computational models benefit from offering a relatively inexpensive method to develop loading models and from their ability to
implicitly incorporate first principles into their analysis, avoiding the pitfall of neglecting the dynamic nature of filtration.
Furthermore, they offer an ability to investigate the loading process on the microscale, allowing insight into processes that are
prohibitively difficult or even impossible to analyze using experimental or analytical techniques.
• Experimental models benefit from their inherent ability to correlate data, allowing for the quantification of difficult processes and
the avoidance of simplifying assumptions that potentially add implicit error. However, this is tempered by the costs associated with
experimentation and the difficulty empirical models experience in addressing encountered anomalies.

Future work regarding particle loading models for fibrous air filters may take several different forms, but two specific avenues are
highlighted here. First, the lack of available data and reliable methods of the measurement of media characteristics such as solidity and
the fiber size distributions lead to the use of assumptions and nominal published values from manufacturers, if available. This
potentially results in avoidable deviations between model predictions and experimental data. Methods useful in characterizing the
filtration media will assist in the pairing of predictions with experimental data, resulting in less confounding factors manifesting in
deviations between predictions and experimental results. Second, most loading models tend to focus on either depth loading or surface
loading, with very little research focusing on the entire loading process. Although arguments may be made that in many instances
surface loading dominates the life cycle of a filter, especially for high-efficiency filters, it is not conceptually or functionally correct to
ignore one of more of the loading regimes. This may potentially yield erroneous predictions in the performance of a filter, although the
amount of error is difficult to express. Research focusing on the transition loading regime will assist in addressing this research gap and
allow for the application of all three loading regimes, covering the full loading process.

Author contributions

Conceptualization, G.B., H.C.; Original draft, G.B., I.B.; Content review and editing, G.B, I.B, H.C.; Project oversight, H.C.; Funding,
H.C.

Funding

This material is based upon work supported by the U.S. Department of Energy Office of Environmental Management under Award
Number: DE-EM0003163.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that have influenced the work
reported in this paper.

30
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

Data availability

No data was used for the research described in the article.

Nomenclature

Term Description Units

A Filtration area m2
Cc Cunningham correction factor Dimensionless
Cd Correction factor for efficiency by diffusion Dimensionless
Cr Correction factor for efficiency by interception Dimensionless
d Diameter m
da Aerodynamic diameter m
df Fiber diameter m
dv Volume equivalent diameter m
dvg Geometric mean diameter of dv m
dg Geometric mean diameter m
dp Particle diameter m
dpf Newly formed particle fiber diameter m
h Adhesion probability Dimensionless
J Layer number Dimensionless
K Specific resistance Pa•s•m•kg-1 or Pa•s•m-1
Knf Knudsen number relative to a fiber Dimensionless
Kk Kozeny constant Dimensionless
Ku Kuwabara hydrodynamic factor Dimensionless
m Mass kg
MMD Mass median diameter m
MPPS Most penetrating particle size m
n Particle concentration m -3
np Number of layers in media Dimensionless
P Penetration Dimensionless
Pe Peclet number Dimensionless
Pe’’ Modified Peclet number Dimensionless
Q Volumetric flow rate m3/s
R Interception parameter Dimensionless
R’ Modified interception parameter Dimensionless
RN Resistance factor, correlation coefficient Dimensionless
r Fiber radius m
St Stokes number Dimensionless
t Time or thickness s or m
U0 Free-stream velocity, gas velocity m/s
VTS Particle settling velocity m/s
v Velocity m/s
wT Surface loading transition point Kg/m2
x Distance into medium m
Z Layer thickness Dimensionless
ΔP Airflow Resistance Pa
ΔPp Particle cake airflow resistance Pa
ΔP0 Clean filter airflow resistance Pa
α Solidity or packing density Dimensionless
αf Solidity or packing density of the fibrous media Dimensionless
αp Solidity or packing density of the particle bed Dimensionless
ε Porosity Dimensionless
η Capture efficiency Dimensionless
ηdiff Diffusion capture efficiency Dimensionless
ηdr Enhance interception efficiency from diffusion Dimensionless
ηelec Electrostatic capture efficiency Dimensionless
ηimp Impaction capture efficiency Dimensionless
ηinter Interception capture efficiency Dimensionless
ηgrav Gravitational settling capture efficiency Dimensionless
ηΣ Total single-fiber efficiency Dimensionless
ρ Density kg/m3
ρp Particle density kg/m3
ρpd Particle deposit density kg/m3
μ Viscosity Pa*s
φ Collision efficiency Dimensionless
σ Standard deviation Dimensionless
σf Standard deviation of fibers Dimensionless
σg Geometric standard deviation Dimensionless
θ Constant denoting loading regime Dimensionless
ζ Mean penetration factor Dimensionless

31
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

References

Abdolghader, P., Brochot, C., Haghighat, F., & Bahloul, A. (2018). Airborne nanoparticles filtration performance of fibrous media: A review. Science and Technology for
the Built Environment, 24(6), 648–672. https://doi.org/10.1080/23744731.2018.1452454
Agay-Shay, K., Friger, M., Linn, S., Peled, A., Amitai, Y., & Peretz, C. (2013). Air pollution and congenital heart defects. Environmental Research, 124, 28–34. https://
doi.org/10.1016/j.envres.2013.03.005
Aguiar, M. L., & Coury, J. R. (1996). cake formation in fabric filtration of gases. Industrial & Engineering Chemistry Research, 35(10), 3673–3679. https://doi.org/
10.1021/ie960042p
Alilou, Y. (2018). Impact sur le colmatage en régimes transitoire et permanent des écoulements d’air induits par le plissage des médias filtrants THE [Doctoral Thesis].
Université de Lorraine.
Angle, D., Banks, E., Bergman, W., Boyce, W., Branagan, E., Brinkley, R., et al. (2003). In J. W. Slawski, J. K. Fretthold, M. R. Hargan, & R. W. Zavadoski (Eds.), DOE
nuclear air cleaning Handbook. U.S. Department of Energy. https://www.standards.doe.gov/standards-documents/1100/1169-bhdbk-2003-pt1/@@images/file.
Artous, S., Bouilloux, L., Marchal, P., & Ouf, F. X. (2012). Clogging of mini-pleat and plane air filters by graphite nanoparticles aggregates simulating combustion aerosol:
Experimental and theoretical description of filter pressure drop and cake porosity. https://www.researchgate.net/profile/Sebastien-Artous-2/publication/268746108_
Clogging_of_mini-pleat_and_plane_air_filters_by_graphite_nanoparticles_aggregates_simulating_fire_combustion_aerosol_experimental_and_theoretical_description_
of_filter_pressure_.
ASME. (2017). ASME AG-1-2017 Code on nuclear air and gas treatment. http://cstools.asme.org/.
Azimian, M., Cheng, L., & Wiegmann, A. (2020). Filtration modeling and simulation with GeoDict®, from filter media to filter element. Math2Market GmbH. https://doi.
org/10.30423/report.m2m-2020-03
Bai, H., Qian, X., Fan, J., Qian, Y., Liu, Y., Duo, Y., et al. (2020). Micro-scale layered structural filtration efficiency model_ Probing filtration properties of non-uniform
fibrous filter media. Separation and Purification Technology, 9.
Becker, J., Cheng, L., Kronsbein, C., & Wiegmann, A. (2016). Simulation of cake filtration for polydisperse particles. Chemical Engineering & Technology, 39(3),
559–566. https://doi.org/10.1002/ceat.201500350
Beckman, I. P., Berry, G., Cho, H., & Riveros, G. (2021). Digital twin geometry for fibrous air filtration media. Fibers, 9.
Bémer, D., & Callé, S. (2000). Evolution of the efficiency and pressure drop of a filter media with loading. Aerosol Science and Technology, 33(5), 427–439. https://doi.
org/10.1080/02786820050204673
Bergman, W. (2006). HEPA filter particle loading. 29th nuclear air cleaning conference. Cincinnati, OH https://www.isnatt.org/.
Bergman, W., Biermann, A. H., Hebard, H. D., Lum, B. Y., & Kuhl, W. D. (1981). Electrostatic air filters generated by electric fields. https://doi.org/10.2172/6391664
Bergman, W., Biermann, A., Kuhl, W., Lum, B., Bogdanoff, A., Hebard, H., et al. (1983). Electric air filtration: Theory , laboratory studies , hardware development , and field
evaluation ; abstract executive siramary present filter-test methods introduction present filter-test methods based on integrated measurements of heterodisperse aeros.
Bergman, W., Taylor, R. D., Miller, H. H., Biermann, A. H., & Hebard, H. D. (1979). Enhanced filtration program at LLL - a progress report (Vol. 1).
Berry, G., Parsons, A., Morgan, M., Rickert, J., & Cho, H. (2022). A review of methods to reduce the probability of the airborne spread of COVID-19 in ventilation
systems and enclosed spaces. Environmental Research, 203, Article 111765. https://doi.org/10.1016/j.envres.2021.111765
Bourrous, S., Bouilloux, L., Nerisson, P., Thomas, D., & Appert-Collin, J. C. (2017). Influence of pleat deformation on pressure drop for a high-efficiency particulate air
filter: A small-scale experimental approach. Journal of Nuclear Engineering and Radiation Science, 3(1), 1–6. https://doi.org/10.1115/1.4034711
Bourrous, S., Bouilloux, L., Ouf, F. X., Appert-Collin, J. C., Thomas, D., Tampère, L., et al. (2014). Measurement of the nanoparticles distribution in flat and pleated
filters during clogging. Aerosol Science and Technology, 48(4), 392–400. https://doi.org/10.1080/02786826.2013.878453
Bourrous, S., Bouilloux, L., Ouf, F.-X., Lemaitre, P., Nerisson, P., Thomas, D., et al. (2016). Measurement and modeling of pressure drop of HEPA filters clogged with
ultrafine particles. Powder Technology, 289, 109–117. https://doi.org/10.1016/j.powtec.2015.11.020
Brach, R. M., & Dunn, P. F. (1998). Models of rebound and capture for oblique microparticle impacts. Aerosol Science and Technology, 29(5), 379–388. https://doi.org/
10.1080/02786829808965577
Brown, R. C. (1993). Air filtration: An integrated approach to the theory and application of fibrous filters. Pergamon Press.
Brown, R. C., & Thorpe, A. (2001). Glass-fibre filters with bimodal fibre size distributions. Powder Technology, 118(1–2), 3–9. https://doi.org/10.1016/S0032-5910
(01)00288-1
Cheng, Y.-H., & Tsai, C.-J. (1998). Factors influencing pressure drop through a dust cake during filtration. Aerosol Science and Technology, 29(4), 315–328. https://doi.
org/10.1080/02786829808965572
Chen, Y. S., & Hsiau, S. S. (2009). Cake formation and growth in cake filtration. Powder Technology, 192(2), 217–224. https://doi.org/10.1016/j.powtec.2008.12.014
Coronas, M. V., Pereira, T. S., Rocha, J. A. V., Lemos, A. T., Fachel, J. M. G., Salvadori, D. M. F., et al. (2009). Genetic biomonitoring of an urban population exposed to
mutagenic airborne pollutants. Environment International, 35(7), 1023–1029. https://doi.org/10.1016/j.envint.2009.05.001
Dahneke, B. (1971). The capture of aerosol particles by surfaces. Journal of Colloid and Interface Science, 37(2), 342–353. https://doi.org/10.1016/0021-9797(71)
90302-X
Davies, C. N. (1953). The separation of airborne dust and particles. Proceedings of the Institution of Mechanical Engineers - Part B: Management and Engineering
Manufacture, 1(1–12), 185–213. https://doi.org/10.1177/095440545300100113
Davies, C. N. (1973). Air filtration. Academic Press Inc.
Deshpande, R., Antonyuk, S., & Iliev, O. (2020). DEM-CFD study of the filter cake formation process due to non-spherical particles. Particuology, 53, 48–57. https://
doi.org/10.1016/j.partic.2020.01.003
Dhaniyala, S., & Liu, B. Y. H. (2001a). Experimental investigation of local efficiency variation in fibrous filters. Aerosol Science and Technology, 34(2), 161–169.
https://doi.org/10.1080/027868201300034745
Dhaniyala, S., & Liu, B. Y. H. (2001b). Theoretical modeling of filtration by nonuniform fibrous filters. Aerosol Science and Technology, 34(2), 170–178. https://doi.
org/10.1080/027868201300034763
Dunn, P. F., & Renken, K. J. (1987). Impaction of solid aerosol particles on fine wires. Aerosol Science and Technology, 7(1), 97–107. https://doi.org/10.1080/
02786828708959150
Emi, H., Okuyama, K., & Adachi, M. (1977). The effect of neighbouring fibers on the single fiber inertia-interception efficiency of aerosols. Journal of Chemical
Engineering of Japan, 10(2), 148–153. https://doi.org/10.1252/jcej.10.148
Endo, Y., Chen, D.-R., & Pui, D. Y. H. (1998a). Bimodal aerosol loading and dust cake formation on air filters. Filtration & Separation, 35(2), 191–195. https://doi.org/
10.1016/S0015-1882(98)91369-6
Endo, Y., Chen, D.-R., & Pui, D. Y. H. (1998b). Effects of particle polydispersity and shape factor during dust cake loading on air filters. Powder Technology, 98(3),
241–249. https://doi.org/10.1016/S0032-5910(98)00063-1
Ergun, S., & Orning, A. A. (1949). Fluid flow through randomly packed columns and fluidized beds. Industrial and Engineering Chemistry, 41(6), 1179–1184. https://
doi.org/10.1021/ie50474a011
Feng, Z., Pan, W., Wang, Y., & Long, Z. (2018). Modeling filtration performance of pleated fibrous filters by Eulerian-Markov method. Powder Technology, 340,
502–510. https://doi.org/10.1016/j.powtec.2018.09.037
Fotovati, S., Hosseini, S. A., Vahedi Tafreshi, H., & Pourdeyhimi, B. (2011). Modeling instantaneous pressure drop of pleated thin filter media during dust loading.
Chemical Engineering Science, 66(18), 4036–4046. https://doi.org/10.1016/j.ces.2011.05.038

32
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

Fotovati, S., Tafreshi, H. V., & Pourdeyhimi, B. (2012). A macroscale model for simulating pressure drop and collection efficiency of pleated filters over time.
Separation and Purification Technology, 98, 344–355. https://doi.org/10.1016/j.seppur.2012.07.009
Fotovati, S., Vahedi Tafreshi, H., & Pourdeyhimi, B. (2010). Influence of fiber orientation distribution on performance of aerosol filtration media. Chemical Engineering
Science, 65(18), 5285–5293. https://doi.org/10.1016/j.ces.2010.06.032
GeoDict. (2019). The Digital Material Laboratory. https://www.math2market.com/showroom/whitepaper.html.
Godoy, C., & Thomas, D. (2020). Influence of relative humidity on HEPA filters during and after loading with soot particles. Aerosol Science and Technology. https://
doi.org/10.1080/02786826.2020.1726278, 0(0), 000.
Happel, J. (1959). Viscous flow relative to arrays of cylinders. AIChE Journal, 5(2), 174–177. https://doi.org/10.1002/aic.690050211
Hasimoto, H. (1959). On the periodic fundamental solutions of the Stokes equations and their application to viscous flow past a cubic array of spheres. Journal of Fluid
Mechanics, 5. https://doi.org/10.1017/S0022112059000222 (02), 317–317.
Henry, F.s., & Ariman, T. (1983). An evaluation of the Kuwabara model. Particulate Science and Technology, 1(1), 1–20. https://doi.org/10.1080/02726358308906350
Hinds, W. C. (1999). Aerosol technology: Properties, behavior, and measurement of airborne particles (2nd ed.). John Wiley & Sons, Inc.
Hinds, W. C., & Kadrichu, N. P. (1997). The effect of dust loading on penetration and resistance of glass fiber filters. Aerosol Science and Technology, 27(2), 162–173.
https://doi.org/10.1080/02786829708965464
Hoch, D., Azimian, M., Baumann, A., Behringer, J., & Niessner, J. (2020). Comparison of voxel-based and mesh-based CFD models for aerosol deposition on complex
fibrous filters. Chemical Engineering & Technology, 43(12), 2538–2547. https://doi.org/10.1002/ceat.202000318
Hosseini, S. A., & Tafreshi, H. V. (2010). Modeling particle filtration in disordered 2-D domains: A comparison with cell models. Separation and Purification Technology,
74(2), 160–169. https://doi.org/10.1016/j.seppur.2010.06.001
Humphries, W. (1980). An investigation of the distribution of fluid velocities in fibrous filters. Powder Technology, 25(1), 31–36. https://doi.org/10.1016/0032-5910
(80)87005-7
Hu, C., Wang, M., Wang, J., Zheng, G., Guo, S., & Lai, C.-H. (2019). Characteristic parameters affecting the filtration performance in fibrous porous media. https://doi.org/
10.1109/DCABES48411.2019.00058
Japuntich, D. A. (1991). CLogging of fibrous filters with monodisperse aerosols, 264–264.
Joubert, A., Laborde, J. C., Bouilloux, L., Chazelet, S., & Thomas, D. (2011). Modelling the pressure drop across HEPA filters during cake filtration in the presence of
humidity. Chemical Engineering Journal, 166(2), 616–623. https://doi.org/10.1016/j.cej.2010.11.033
Jung, L. (1987). Impact of air-filter condition on HVAC (heating, ventilation, and air conditioning) equipment.
Kanaoka, C., Emi, H., Hiragi, S., & Myojo, T. (1986). Morphology of particulate agglomerates on a cylindrical fiber and collection efficiency of a dust loaded fiber. In
Aerosols—formation and reactivity (proceedings of the 2nd international aerosol conference (pp. 674–677). Belin: Pergamon JOurnals Ltd, 1986.
Kanaoka, C., & Hiragi, S. (1990). Pressure drop of air filter with dust load. Journal of Aerosol Science, 21(1), 127–137. https://doi.org/10.1016/0021-8502(90)90027-
U
Kasper, G., Schollmeier, S., & Meyer, J. (2010). Structure and density of deposits formed on filter fibers by inertial particle deposition and bounce. Journal of Aerosol
Science, 41(12), 1167–1182. https://doi.org/10.1016/j.jaerosci.2010.08.006
Kasper, G., Schollmeier, S., Meyer, J., & Hoferer, J. (2009). The collection efficiency of a particle-loaded single filter fiber. Journal of Aerosol Science, 40(12),
993–1009. https://doi.org/10.1016/j.jaerosci.2009.09.005
Keuken, M., Zandveld, P., van den Elshout, S., Janssen, N. A. H., & Hoek, G. (2011). Air quality and health impact of PM10 and EC in the city of Rotterdam, The
Netherlands in 1985–2008. Atmospheric Environment, 45(30), 5294–5301. https://doi.org/10.1016/j.atmosenv.2011.06.058
Kim, K.-H., Kabir, E., & Kabir, S. (2015). A review on the human health impact of airborne particulate matter. Environment International, 74, 136–143. https://doi.org/
10.1016/j.envint.2014.10.005
Kim, S. C., Wang, J., Shin, W. G., Scheckman, J. H., & Pui, D. Y. H. (2009). Structural properties and filter loading characteristics of soot agglomerates. Aerosol Science
and Technology, 43(10), 1033–1041. https://doi.org/10.1080/02786820903131081
Koch, M., & Krammer, G. (2008). The permeability distribution (PD) method for filter media characterization. Aerosol Science and Technology, 42(6), 433–444. https://
doi.org/10.1080/02786820802172053
Kuwabara, S. (1959). The forces experienced by randomly distributed parallel circular cylinders or spheres in a viscous flow at small Reynolds numbers. Journal of the
Physical Society of Japan, 14(4), 527–532. https://doi.org/10.1143/JPSJ.14.527
Laborde, J. C., Fabbro, L. D. E. L., Mocho, V. M., & Ricciardi, L. (2002). Contribution to the modelling of industrial pleated filters by solid particles.
Lee, K., Jung, Y.-W., Park, H., Kim, D., & Kim, J. (2021). Sequential multiscale simulation of a filtering facepiece for prediction of filtration efficiency and resistance in
varied particulate scenarios. ACS Applied Materials & Interfaces, 13(48), 57908–57920. https://doi.org/10.1021/acsami.1c16850
Lee, K. W., & Liu, B. Y. H. (1980). On the minimum efficiency and the most penetrating particle size for fibrous filters. Journal of the Air Pollution Control Association, 30
(4), 377–381. https://doi.org/10.1080/00022470.1980.10464592
Lee, K. W., & Liu, B. Y. H. (1981). Experimental study of aerosol filtration by fibrous filters. Aerosol Science and Technology, 1(1), 35–46. https://doi.org/10.1080/
02786828208958577
Lee, K. W., & Liu, B. Y. H. (1982). Theoretical study of aerosol filtration by fibrous filters. Aerosol Science and Technology, 1(2), 147–161. https://doi.org/10.1080/
02786828208958584
Lee, K. W., & Ramamurthi, M. (2001). Filter collection. In P. A. Baron, & K. Willeke (Eds.), Aerosol measurement: Principles, techniques, and applications (second (pp.
197–228). John Wiley & Sons, Inc.
Letourneau, P., Mulcey, P., & Vendel, J. (1990). Aerosol penetration inside HEPA filtration media.
Letourneau, P., Renaudin, V., & Vendel, J. (1992). Effects of the particle penetration inside the filter medium on the HEPA filter pressure drop. In Nuclear air cleaning
and treatment conference. Denver, CO https://inis.iaea.org/search/search.aspx?orig_q=RN:24072464.
Leung, W. W.-F., & Hung, C.-H. (2008). Investigation on pressure drop evolution of fibrous filter operating in aerodynamic slip regime under continuous loading of
sub-micron aerosols. Separation and Purification Technology, 63(3), 691–700. https://doi.org/10.1016/j.seppur.2008.07.015
Linden, S., Cheng, L., & Wiegmann, A. (2018). Specialized methods for direct numerical simulations in porous media. Math2Market GmbH. https://doi.org/10.30423/
report.m2m-2018-01
Liu, M., Claridge, D. E., & Deng, S. (2003). An air filter pressure loss model for fan energy calculation in air handling units. International Journal of Energy Research, 27
(6), 589–600. https://doi.org/10.1002/er.897
Maddineni, A. K., Das, D., & Damodaran, R. M. (2018). Air-borne particle capture by fibrous filter media under collision effect: A CFD-based approach. Separation and
Purification Technology, 193, 1–10. https://doi.org/10.1016/j.seppur.2017.10.065
Math2Market. Your partner for innovative material development. https://www.math2market.com/math2market/about-us.html. (n.d.). (Accessed 22 July 2022.).
Mukherjee, A., & Agrawal, M. (2017). World air particulate matter: Sources, distribution and health effects. Environmental Chemistry Letters, 15(2), 283–309. https://
doi.org/10.1007/s10311-017-0611-9
Novick, V. J., Higgins, P. J., Dierkschiede, D., Abrahamson, C., & Richardson, W. B. (1991). Efficiency and mass loading Characteristics of a typical HEPA filter media
material. 782–798. https://inis.iaea.org/search/search.aspx?orig_q=RN:22085832.
Novick, V. J., Klassen, J. F., & Monson, P. R. (1992, August). Predicting mass loading as a function of pressure difference across prefilter/HEPA filter systems. In 22nd DOE/
NRC nuclear air cleaning and treatment conference. Denver, CO https://www.osti.gov/biblio/7047281.
Novick, V. J., Monson, P. R., & Ellison, P. E. (1992). The effect of solid particle mass loading on the pressure drop of HEPA filters. Journal of Aerosol Science, 23(6),
657–665. https://doi.org/10.1016/0021-8502(92)90032-Q
Paw U, K. T. (1984). Dimensional aspects of aerosol deposition on cylinders with rebound. Journal of Aerosol Science, 15(6), 657–660. https://doi.org/10.1016/0021-
8502(84)90003-X
Payatakes, A. C. (1976). Model of the dynamic behavior of a fibrous filter. Application to case of pure interception during period of unhindered growth. Powder
Technology, 14(2), 267–278. https://doi.org/10.1016/0032-5910(76)80075-7

33
G. Berry et al. Journal of Aerosol Science 167 (2023) 106078

Payatakes, A. C. (1977). Model of transient aerosol particle deposition in fibrous media with dendritic pattern. AIChE Journal, 23(2), 192–202. https://doi.org/
10.1002/aic.690230208
Payatakes, A. C., & Gradoń, L. (1980). Dendritic deposition of aerosol particles in fibrous media by inertial impaction and interception. Chemical Engineering Science,
35(5), 1083–1096. https://doi.org/10.1016/0009-2509(80)85097-4
Payatakes, A. C., & Okuyama, K. (1982). Effects of aerosol particle deposition on the dynamic behavior of uniform or multilayer fibrous filters. Journal of Colloid and
Interface Science, 88(1), 55–78. https://doi.org/10.1016/0021-9797(82)90155-2
Payet, S., Boulaud, D., Madelaine, G., & Renoux, A. (1992). Penetration and pressure drop of a HEPA filter during loading with submicron liquid particles. Journal of
Aerosol Science, 23(7), 723–735. https://doi.org/10.1016/0021-8502(92)90039-X
Penicot, P., Thomas, D., Contal, P., Leclerc, D., & Vendel, J. (1999). Clogging of HEPA fibrous filters by solid and liquid aerosol particles: An experimental study.
Filtration & Separation, 36(2), 59–64. https://doi.org/10.1016/S0015-1882(99)80036-6
Podgórski, A. (2009). Estimation of the upper limit of aerosol nanoparticles penetration through inhomogeneous fibrous filters. Journal of Nanoparticle Research, 11(1),
197–207. https://doi.org/10.1007/s11051-008-9472-2
Podgorski, A., Maißer, A., Szymanski, W. W., Jackiewicz, A., & Gradon, L. (2011). Penetration of monodisperse, singly charged nanoparticles through polydisperse
fibrous filters. Aerosol Science and Technology, 45(2), 215–233. https://doi.org/10.1080/02786826.2010.531300
Rao, N., & Faghri, M. (1988). Computer modeling of aerosol filtration by fibrous filters. Aerosol Science and Technology, 8(2), 133–156. https://doi.org/10.1080/
02786828808959178
Saleh, A. M., Fotovati, S., Vahedi Tafreshi, H., & Pourdeyhimi, B. (2014). Modeling service life of pleated filters exposed to poly-dispersed aerosols. Powder Technology,
266, 79–89. https://doi.org/10.1016/j.powtec.2014.06.011
Schweers, E., & Löffler, F. (1994). Realistic modelling of the behaviour of fibrous filters through consideration of filter structure. Powder Technology, 80(3), 191–206.
https://doi.org/10.1016/0032-5910(94)02850-8
Shou, D., Fan, J., Zhang, H., Qian, X., & Ye, L. (2015). Filtration efficiency of non-uniform fibrous filters. Aerosol Science and Technology, 49(10), 912–919. https://doi.
org/10.1080/02786826.2015.1083092
Shu, Z., Qian, F., Zhu, J., & Lu, J. (2020). Numerical simulation of gas-solid flow characteristic of particles in fibrous media using OpenFOAM. Indoor and Built
Environment, 29(7), 921–930. https://doi.org/10.1177/1420326X19862016
Song, C. B., Lee, J. L., Park, H. S., & Lee, K. W. (2007). Effect of solid monodisperse particles on the pressure drop of fibrous filters. Korean Journal of Chemical
Engineering, 24(1), 148–153. https://doi.org/10.1007/s11814-007-5024-1
Song, C. B., & Park, H. S. (2006). Analytic solutions for filtration of polydisperse aerosols in fibrous filter. Powder Technology, 170(2), 64–70. https://doi.org/10.1016/
j.powtec.2006.08.011
Spielman, L., & Goren, S. L. (1968). Model for predicting pressure drop and filtration efficiency in fibrous media. Environmental Science and Technology, 2(4), 279–287.
https://doi.org/10.1021/es60016a003
Stechkina, I. B., Kirsch, A. A., & Fuchs, N. A. (1969). Studies on fibrous aerosol filters-iv calculation of aerosol deposition in model filters in the range of maximum
penetration. Annals of Occupational Hygiene, 12(1), 1–8. https://doi.org/10.1093/annhyg/12.1.1
Sun, Z., Liang, Y., He, W., Jiang, F., Song, Q., Tang, M., et al. (2019). Filtration performance and loading capacity of nano-structured composite filter media for
applications with high soot concentrations. Separation and Purification Technology, 221(March), 175–182. https://doi.org/10.1016/j.seppur.2019.03.087
Tahir, M. A., & Tafreshi, H. V. (2009). Influence of fiber orientation on the transverse permeability of fibrous media. Physics of Fluids, 21(8). https://doi.org/10.1063/
1.3211192, 083604–083604.
Théron, F., Joubert, A., & Le Coq, L. (2017). Numerical and experimental investigations of the influence of the pleat geometry on the pressure drop and velocity field
of a pleated fibrous filter. Separation and Purification Technology, 182, 69–77. https://doi.org/10.1016/j.seppur.2017.02.034
Thomas, D. (2017). 5—filtration of solid aerosols. In Aerosol filtration (1st ed., pp. 123–159). Elsevier.
Thomas, D., Contal, P., Renaudin, V., Penicot, P., Leclerc, D., & Vendel, J. (1999). Modelling pressure drop in hepa filters during dynamic filtration. Journal of Aerosol
Science, 30(2), 235–246. https://doi.org/10.1016/S0021-8502(98)00036-6
Thomas, D., Penicot, P., Contal, P., Leclerc, D., & Vendel, J. (2001). Clogging of fibrous filters by solid aerosol particles Experimental and modelling study. Chemical
Engineering Science, 56(11), 3549–3561. https://doi.org/10.1016/S0009-2509(01)00041-0
Tien, C., & Ramarao, B. V. (2013). Can filter cake porosity be estimated based on the Kozeny–Carman equation? Powder Technology, 237, 233–240. https://doi.org/
10.1016/j.powtec.2012.09.031
Wang, C., & Otani, Y. (2013). Removal of nanoparticles from gas streams by fibrous filters: A review. Industrial & Engineering Chemistry Research, 52(1), 5–17. https://
doi.org/10.1021/ie300574m
Xiao, B., Zhang, Y., Wang, Y., Jiang, G., Liang, M., Chen, X., et al. (2019). A fractal model for kozeny–carman constant and dimensionless permeability of fibrous
porous media with roughened surfaces. Fractals, 27. https://doi.org/10.1142/S0218348X19501160 (07), 1950116–1950116.
Yang, H., He, S., Ouyang, H., Anderson, M. J., Shen, L., & Hogan, C. J. (2018). The pressure drop across combined polydisperse spherical particle – cylindrical fiber
networks. Chemical Engineering Science, 192, 634–641. https://doi.org/10.1016/j.ces.2018.08.006
Yeh, H.-C., & Liu, B. Y. H. (1974a). Aerosol filtration by fibrous filters—I. Theoretical. Journal of Aerosol Science, 5(2), 191–204. https://doi.org/10.1016/0021-8502
(74)90049-4
Yeh, H.-C., & Liu, B. Y. H. (1974b). Aerosol filtration by fibrous filters—II. Experimental. Journal of Aerosol Science, 5(2), 205–217. https://doi.org/10.1016/0021-
8502(74)90050-0

34

You might also like