704
Views
5
CrossRef citations to date
0
Altmetric
Original Articles

The Single-Fiber Collision Rate and Filtration Efficiency for Nanoparticles II: Extension to Arbitrary-Shaped Particles

, &
Pages 886-895 | Received 24 Sep 2013, Accepted 10 Jun 2014, Published online: 11 Aug 2014

Abstract

We extend the equations for the dimensionless collision kernel and filtration efficiency, attained previously via mean first-passage time (MFPT) calculations, to particles of arbitrary shape. Specifically, we show that the regression equations for the dimensionless collision rate found considering particle-fiber collisions driven by simultaneous diffusion and interception remain valid for non-spherical particles, provided that an appropriate collision length scale for the non-spherical particle (L) is defined and incorporated into the definitions of the dimensionless collision rate (H) and the diffusive Knudsen number (KnD). Regression equations are provided to calculate this length scale for quasifractal aggregates of varying fractal dimension, as well as cylinders. MFPT calculations reveal that, over ∼5 orders of magnitude in H, these regression equations for the collision length are valid. Furthermore, using the previously attained proportionality between the predicted dimensionless collision rate and the single-fiber efficiency, comparison is made between the equations presented here and measurements of the penetration of both multiwalled carbon nanotubes and quasifractal aggregates through fibrous filters. Reasonable agreement is found between measured and predicted single-fiber efficiencies in both circumstances, supporting the use of the single-fiber efficiency calculation approach we developed.

Copyright 2014 American Association for Aerosol Research

INTRODUCTION

As discussed in detail in a companion study (Hunt et al. Citation2014), the single-fiber efficiency, which is commonly determined as a means to predict the penetration of particles through fibrous filter media, is linearly proportional to the particle-fiber collision rate. Determination of the particle penetration through a filter can, therefore, be reframed as a collision rate-determination problem, which, as shown in the companion study and elsewhere, can be solved under a variety of circumstances via mean first-passage time (MFPT) calculations (Narsimhan and Ruckenstein Citation1985; Gopalakrishnan and Hogan Citation2011; Gopalakrishnan et al. Citation2013b). Specifically for particle-fiber collisions, MFPT calculations show that the dimensionless collision kernel (H) within a Kuwabara (Citation1959) flow cell model, neglecting particle inertia, is described by the following equation: [1a] [1b] [1c]

where KnD is the diffusive Knudsen number (Dahneke Citation1983; Gopalakrishnan and Hogan Citation2011), R is the particle size to fiber size ratio, and Vf is the filter solid-volume fraction. For spherical particles and cylindrical fibers, H, KnD, R, and χf are defined by the following respective equations: [2a] [2b] [2c] [2d]

where β is the two-dimensional collision rate coefficient/kernel, f is the particle friction factor (Dahneke Citation1973; Zhang et al. Citation2012), mp is the particle mass, k is the Boltzmann's constant, T is the temperature,U0 is the fluid flow velocity upstream of the filter, and ap and af are the particle and fiber radii, respectively. Correspondingly, the penetration (P) of spherical particles through a filter of thickness w can be calculated as: [3]

In an earlier report (Hunt et al. Citation2014), Equation (3) was shown to agree reasonably well with the predictions of traditional depth-filtration theory considering particle deposition by diffusion and interception (Stechkina and Fuchs Citation1966; Kirsch et al. Citation1974; Lee and Liu Citation1982) as well as correlations derived from experimental measurements of the heat transfer from banks of cylinders (Žukauskas Citation1972); therefore, Equation (3) can be used as an alternative to prior methods to estimate nanoparticle penetration through fibrous filter media under conditions where χf1/2/KnD < 104, χf1/2KnD < 0.1, Vf ≤ 0.1, and R ≤ 0.2. However, like in the classical depth filtration theory, Equations (1–3) have been developed considering only spherical particles. High-temperature synthesis in the gas phase routinely leads to the formation of non-spherical nanoparticles composed of a number of point-contacting primary spheres (aggregates/agglomerates; ref. Schmidt-Ott et al. Citation1990; Sorensen Citation2011), which are collected via filtration for later use. Further, non-spherical aggregates are frequently a byproduct of incomplete combustion reactions (Chakrabarty et al. Citation2009; Sorensen et al. Citation2003; Latin et al. Citation2013), and fibrous filtration is a commonly used method to control such particulate combustion emissions. Therefore, calculations of particle penetration applicable to non-spherical particles are of considerable interest.

The purpose of this report is, thus, to extend the MFPT calculation approach to determine the dimensionless particle-fiber collision rate considering non-spherical particles, while continuing to account for particle deposition due to diffusion and interception. This issue has been the focus of both experimental and numerical studies previously. Specifically, Kim et al. (Citation2009) examined experimentally the penetration of quasifractal aggregates of varying morphology through fibrous filters. By measuring particle penetration as a function of “mobility equivalent diameter” (with a differential mobility analyzer [DMA]) they were able to compare the penetrations for particles of differing morphology yet having similar diffusion coefficients. At small mobility diameters (higher diffusion coefficients), particle morphology was observed to minimally influence penetration, while at higher mobility diameters (lower diffusion coefficients) it was observed that aggregates were captured more efficiently. Similar findings have been reported by Seto et al. (Citation2010) and, later, by Wang et al. (CitationWang et al. 2011a,b), and both found that, at larger mobility diameters, carbon nanotubes and nanotube bundles were collected more efficiently than their spherical counterparts. Balazy and Podgorski (Citation2007) used calculations of Brownian Dynamics (with a similar framework to the MFPT approach) to examine aggregate deposition onto fibers, but without explicitly accounting for the aggregate morphology and, instead, assuming particular relationships for aggregate effective radii/collision length scales. In doing so, they also predicted enhanced deposition for aggregates. Collectively, these studies demonstrate that non-spherical particles deposit onto filter fibers more efficiently; however, they also indicate that further investigation into the filtration of non-spherical particles is warranted, with more detailed modeling of particle morphology.

In the subsequent sections, the modifications to the results of the MFPT calculation required to determine the dimensionless collision kernel for non-spherical particles with cylindrical fibers are described. Analysis is performed for quasifractal aggregates composed of point-contacting, monodisperse, primary spheres as well as for cylindrical particles. It is shown that, for these non-spherical particles, Equations (1a–c) remain valid, and it is only the definitions provided in Equations (2a–c) that need to be modified slightly.

THEORETICAL METHODS

Collision Length Scale Calculations for Non-Spherical Particles

As noted by Hunt et al. (Citation2014), as with coagulation and diffusion charging, particle deposition onto filter fibers can be analyzed as a transported limited-collision processes. In examining collisions between a point mass and a non-spherical particle (Gopalakrishnan et al. Citation2011, Citation2013b), as well as two non-spherical particles (Thajudeen et al. Citation2012), it has been found that the dimensionless collision kernel valid for sphere–sphere collisions also applies to collisions involving non-spherical entities, provided the modified collision length scales are incorporated into the dimensionless parameters H and KnD. Here, we similarly define collision length scales for non-spherical particles with circular fibers to recast H, KnD, and R, which enables us to test the hypothesis that Equations (1a–c) remain valid with appropriately modified definitions of these parameters. We note that the analysis differs slightly from prior studies of particle–particle collisions; particle–particle collisions occur in a three-dimensional environment, for which a collision radius (the Smoluchowski radius) can be defined for diffusive collisions (the continuum regime), and a collision-projected area (two-dimensional) can be defined for ballistic collisions (the free molecular regime; ref. Thajudeen et al. Citation2012). Conversely, particle–filter fiber collisions may be treated as two-dimensional when filter fibers are modeled as infinitely long, which leads to the requirement that only a ballistic collision length scale be defined (i.e., diffusive collision rates are not a function of collision length scale in two-dimensional environments).

The ballistic collision length for a particle of arbitrary shape with an infinitely long fiber is defined as the orientationally averaged maximum distance of closest approach at which the particle and fiber are in contact. An outline of the calculation procedure for this collision length scale is provided in , which is the two-dimensional analog to the orientationally averaged collision area calculation approach described by Zurita-Gotor and Rosner (Citation2002) and Thajudeen et al. (2012). We consider a non-spherical particle approaching a cylindrical fiber at a specific orientation and random location. Irrespective of the particle's trajectory, at closest approach, the particle's center of mass will be at a distance af + x from the center of the fiber, and, therefore, it is at a distance x from the edge of the fiber. For the orientation in question, at the distance of closest approach, the particle may or may not be in contact with the fiber, which is readily determined by checking if any portion of the particle crosses the x-axis origin (when the axes are aligned, as is depicted in ). By selecting random particle orientations (in three dimensions) and examining values of x selected from a uniform distribution between zero and a value xmax, the collision length between particle and fiber can be defined as a+ L, where L is calculated as: [4]

FIG. 1. Depiction of the procedure used to calculate the collision length scale L for an arbitrary-shaped particle with a cylindrical filter fiber of radius af. x denotes the distance of a particle's center of mass to the fiber surface when at closest approach. In determining L, random particle orientations and values of x are selected to see if the particle and fiber overlap.

FIG. 1. Depiction of the procedure used to calculate the collision length scale L for an arbitrary-shaped particle with a cylindrical filter fiber of radius af. x denotes the distance of a particle's center of mass to the fiber surface when at closest approach. In determining L, random particle orientations and values of x are selected to see if the particle and fiber overlap.

In Equation (4), Ncol is the number of trials resulting in contact, and Ntot is the total number of trials. Provided that (1) xmax is sufficiently large, such that the particle does not collide with the fiber in any orientation when its distance of closest approach is xmax, and (2) a sufficient number of trials are performed (i.e., more than 5000), such that Ncol/Ntot < 1, then L is independent of both Ntot and xmax. When these criteria are satisfied, af + L is the orientationally averaged collision length scale and, with this collision length defined, we propose that Equations (1a–c) remain valid, with Equations (2a–c) modified as: [5a] [5b] [5c]

We note that, because the definition of χf does not involve a description of particle size and shape, the definition of χf in Equation (2d) remains valid for all particles. From Equation (4), it is evident that L = ap for spherical particles; therefore, Equations (5a–c) are more general versions of Equations (2a–c). Although L can be calculated for any particle with the noted procedure, here we calculate L for quasifractal aggregates composed of point contacting, equal-radius spheres, which approximately obey the relationship: [6]

where Nmon is the number of monomer spheres per aggregate, kf is the pre-exponential factor, Rg is the aggregate's radius of gyration, amon is monomer radius, and Df is the fractal dimension. In doing so, aggregates are produced which satisfy Equation (6) (to within 0.1% error) using a cluster–cluster algorithm described by Fillipov et al. (Citation2000) in the range Nmon = 20–1000, kf = 1.3 and 1.7, and Df = 1.5–2.4. This algorithm is selected because it is found to produce aggregates with structural characteristics (Huang et al. Citation1998) similar to those of similar kf and Df generated by diffusion-limited aggregation algorithms. However, to generate aggregates with Df = 1.3 and 2.7, we opt to use the sequential algorithm (Filippov et al. Citation2000; Thajudeen and Hogan Citation2012); with the cluster–cluster algorithm, it is difficult to generate such nearly linear (Df = 1.3) and highly compact (Df = 2.7) aggregates. Both the cluster–cluster and sequential algorithms are capable of producing random structures, and for each Nmon, kf, and Df examined, 10 aggregates are produced (which are described by the coordinates of their monomers). We similarly calculate L for cylinders of varying radii, acyl, and aspect ratios (Rcyl, the cylinder's radius to length ratio). For both aggregates and cylinders, regression equations (for L/amon for aggregates and for L/acyl for cylinders, respectively) are developed, enabling direct calculation of L from the structural descriptors of the particle in question.

Non-Spherical Particle MFPT Calculations

The validity of Equations (5a–c) is examined via the determination of H from MFPT calculations for quasifractal aggregates. The calculation procedure is represented in . As in the investigation of spherical particle filtration (Hunt et al. Citation2014), MFPT calculations are performed by monitoring a particle's motion in a Kuwabara flow cell using a dimensionless solution to the Langevin equations. For a non-spherical particle, we define the dimensionless velocity vector as and its dimensionless position as , leaving the solution to the equations of motion as Equations (7a–d) of Hunt et al. (Citation2014). Simulations are, therefore, performed similarly to the manner described by Hunt et al., to determine H as a function of input KnD, R (which scales the dimensionless radii of primary particles), χf, and Vf. We, again, restrict simulations to situations to the conditions χf1/2/KnD < 104, χf1/2KnD < 0.1, Vf ≤ 0.1, and R ≤ 0.2. Inertial effects are included in particle motion, although such effects are small for the range of dimensionless input variables examined. The rotational motion of aggregates during their migration through the Kuwabara cell is neglected; based upon the calculations of Zurita-Gotor and Rosner (Citation2002), rotations are not expected to substantially alter the particle–fiber collision rate for values of R in the noted range. The only differences between the MFPT calculation approach for non-spherical particles and spherical particles arise when (1) checking if the particle collided with the central fiber, and (2) re-introducing the particle into the Kuwabara cell in instances where it leaves the cell prior to collision. In (1), with spherical particles, collision was assured when the particle center was a dimensionless distance ≤1 from the center of the cell. Conversely, in the case of aggregates, each portion of the aggregate must to examined to determine whether it is in contact with the central fiber. For example, for aggregates with a dimensionless primary particle radius , if any primary sphere center is at a distance or less from the center of the Kuwabara cell, collision has occurred. In (2), each time an aggregate is introduced into the Kuwabara cell, either to commence a calculation or because it left the cell without collision, it is oriented randomly (about its center of mass, and in three dimensions) via the selection of three Euler angles. MFPT calculations are performed for the eight noted aggregates in . We note that, while L is used in the non-dimensionalization of the equations of motion for aggregates, this choice is arbitrary; any length scale can be used to normalize dimensions. However, if another length scale is selected, computational results must be rescaled to infer the input KnD and R, and the output H (Gopalakrishnan et al. Citation2011).

Table 1. A summary of the quasifractal aggregates used in mean first-passage time calculations

FIG. 2. Depiction of the mean first-passage time calculation approach for aggregates, which may be contrasted with the method of Hunt et al. (Citation2014) for spherical particles.

FIG. 2. Depiction of the mean first-passage time calculation approach for aggregates, which may be contrasted with the method of Hunt et al. (Citation2014) for spherical particles.

RESULTS AND DISCUSSION

Non-Spherical Particle Collision Length Scale and Comparison to MFPT Calculations

Plots of L/amon as a function of Rg/amon for aggregates (averaged over 10 aggregates) and L/acyl as a function of 1/Rcyl are shown in and 3b, respectively. Interestingly, for aggregates, results collapse to curves independent of kf; L/amon versus Rg/amon is nearly linear (R2 > 0.999), with slope dependent on Df in all instances. The linear slope (1/α) linking L/amon to Rg/amon, as determined from calculations, is summarized in . Moreover, combining all results leads to the regression equations: [7a] [7b]

Table 2. A summary of the slopes (1/α) linking L/amon to Rg/amon for quasifractal aggregates

FIG. 3. Plots of the collision length scale to radius ratio as a function of Rg/amon (a) for aggregates and (b) 1/Rcyl for cylinders.

FIG. 3. Plots of the collision length scale to radius ratio as a function of Rg/amon (a) for aggregates and (b) 1/Rcyl for cylinders.

Results for cylinders lead to the regression equation: [7c]

We note that the regression equations for L of aggregates and cylinders gives distinctly different results from the approach of Balazy and Podgorski (Citation2007), who assumed that L = [(2+Df)/Df]1/2Rg, i.e., that L is equal to the approximate outer radius of an aggregate. The outer radius and the values of L calculated here only agree well with one another as Df → 3, and differ by a factor of up to 2 at lower fractal dimensions. Substantial disagreement between L and the outer radius is also found for cylinders, as the outer radius would be defined as half the cylinder length. Therefore, when examining both chain-like aggregates and cylinders, the calculations here are in clear disagreement with the calculations of Balazy and Podgorski (Citation2007).

The predictions of Equations (1a–c), using the definitions in Equation (5a–c) and results of MFPT calculation for quasifractal aggregates, are plotted against one another in . For guidance, a 1:1 line is additionally displayed. Across ∼5 orders of magnitude in H, independent of aggregate shape, equation predictions and MFPT calculation results typically agree to within 10% of one another. This is similar to the agreement found between Equation (1a–c) predictions and MFPT calculations with spheres. Therefore, we propose that Equation (1a–c) can be used with confidence to predict the dimensionless collision rate and, subsequently, the single-fiber efficiency, for non-spherical particles.

FIG. 4. A comparison of H inferred from mean first passage time calculations to the values of H calculated with Equations (1a–c) for quasifractal aggregates, with properties specified in . A 1:1 line is shown for guidance.

FIG. 4. A comparison of H inferred from mean first passage time calculations to the values of H calculated with Equations (1a–c) for quasifractal aggregates, with properties specified in Table 1. A 1:1 line is shown for guidance.

Comparison to Experimental Measurements

We now use both Equation (1a–c) predictions and MFPT simulations to compare them to the experimental measurements of the penetration of multiwalled carbon nanotubes through fibrous filters reported by Seto et al. (Citation2010). These penetration experiments were performed using multiwalled nanotubes with three different radii and aspect ratios, characterized by electron microscopy. A summary of the comparison is provided in , wherein calculations of the parameters χf, KnD, Vf, R, and H are based on the reported particle penetrations, face velocities, the filter volume fraction (0.049), and average filter-fiber radius (1.4 μm). To define KnD, it is necessary to calculate the mass (although in the absence of inertia, the mass does not influence results) as well as the friction factor of the particles under examination. Modeling nanotubes as straight cylinders, the mass can be calculated from knowledge of the nanotube material density, radius, and aspect ratio. Particle friction factors can be inferred through use of a DMA; however, while Seto et al. (Citation2010) used a DMA upstream of their filtration system to classify particles based on their electrical mobilities (charge to friction factor ratios), many particles transmitted through the DMA may have been multiply charged (Gopalakrishnan et al. Citation2013a) and, for nanotubes with supermicrometer lengths, particles may have aligned during migration in an electrostatic field (Kasper and Shaw Citation1982). With the assumption that the nanotube does not align preferentially in any direction as it migrates through the filter medium, its friction factor (and similarly the friction factor of any particle) can be calculated considering non-continuum drag effects as (Dahneke Citation1973; Zhang et al. Citation2012): [8a]

Table 3. A summary of the comparison of equation predictions to the measurements of Seto et al. (Citation2010). The noted length and acyl values correspond to the lengths and radii of the nanotubes studied; R, KnD, and χf are the calculated values based upon the reported nanotube and filter fiber dimensions, and U0 is the reported filter-face velocity. Ef (exp.) refers to experimentally measured values, while Ef (pred.) refers to Equation (9) predictions made either with Equations (1a–c) or with MFPT calculation results. The percentage difference (% Diff.) between experiments and calculations is normalized by experimental single-fiber efficiencies

where μ is background gas dynamic viscosity, λ is the gas molecule hard sphere mean free path, PA is the particle's orientationally averaged projected area, and RS is the particle's Smoluchowski radius (Gopalakrishnan et al. Citation2011), equated with its hydrodynamic radius via the Hubbard–Douglas approximation (Douglas et al. Citation1994). Shown previously (Hansen Citation2004), for cylinders RS and PA are given by the following equations: [8b] [8c]

Therefore, for all cylinders under investigation, L, Rs, and PA, and, therefore, KnD, R, and χf (length scale independent) are calculable, as is H by either the regression equations or from MFPT simulations directly. Following the prior report (Hunt et al. 2014), the single-fiber efficiency, Ef, is related to H via the equation: [9]

Values of Ef inferred from (1) the measurements of Seto et al. (Citation2010), (2) extrapolations of the regression Equations (1a–c), and (3) directly from MFPT simulations using the input KnD, χf, Vf, and R corresponding to the mean values for each carbon nanotube sample reported by Seto et al. are displayed in . In addition, for each of the three mobility-classified nanotubes examined, these values are plotted as a function of χf1/2/KnD in , with predictions of Equation (9) shown for R = 0.26, 0.31, and 0.49 (the inferred values for the provided nanotube and filter-fiber dimensions), respectively. Results are plotted in this manner because single-fiber efficiencies are solely a function of χf1/2/KnD, which is proportional to the Peclet number (Hunt et al. 2014), once R and Vf are specified.

FIG. 5. A comparison of predicted (dashed lines for regression predictions and closed symbols for MFPT results) and measured (open symbols) single-fiber efficiencies as a function of χf1/2/KnD for multiwalled carbon nanotube deposition on a fibrous filter. Measurement results are taken from Seto et al. (Citation2010).

FIG. 5. A comparison of predicted (dashed lines for regression predictions and closed symbols for MFPT results) and measured (open symbols) single-fiber efficiencies as a function of χf1/2/KnD for multiwalled carbon nanotube deposition on a fibrous filter. Measurement results are taken from Seto et al. (Citation2010).

Furthermore, although disparities between measurements, regression-based calculations, and MFPT results are certainly evident for all three nanotube types, qualitatively, calculations agree well with measurements, and agree well with measurements quantitatively in a number of circumstances (e.g., for R = 0.26 measurements and MFPT calculations are within 10% of one another). Given that none of the parameters input into calculations were “fit” to force agreement with measurements, and simultaneously that a number of features of experiments were not considered in calculations, we believe this comparison supports the applicability of the presented equations for non-spherical particle filtration modeling. Specifically, not included in calculations are the polydispersities in nanotube diameters and lengths, which are non-negligible even downstream of a DMA, the possible preferential orientations of nanotubes as they migrate through the filter, filter-fiber radius, and volume fraction polydispersities, as well as possible electrostatic influences on nanotube motion. Unfortunately, each of these influences is rather involved to consider in detail; we are aware of no study (theoretical or experimental) of non-spherical particle filtration that has considered all such effects in modeling filtration.

For additional comparison of calculations to measurements, we also utilize the results of Kim et al. (Citation2009), who studied the penetration of soot aggregates through fibrous filters. Aggregates were characterized through tandem DMA-aerosol particle mass analyzer (APM; Tajima et al. Citation2011), yielding the effective mass-mobility exponent (Dmm) for examined particles (Scheckman et al. Citation2009). However, while the mass-mobility exponent provides some description of particle morphology, it is not equivalent to the fractal dimension of particles (Sorensen Citation2011), and for particles with mobility equivalent diameters in the 10–500 nm size range at room temperature and atmospheric pressure, a link between the mass-mobility exponent and fractal dimension has not yet been clearly established. Therefore, as with the comparison to nanotube penetration data, there is considerable ambiguity into the true morphology of the particles under examination. With this issue in mind, we make comparison to soot aggregate measurements as follows: DMA and APM measurements yielded the aggregate friction factor (quantified by the mobility equivalent diameter, dm, which was selected to be 30, 50, 100, 150, 200, 250, and 300 nm) and mass, respectively, and, through electron microscopy, Kim et al. found the average filter-fiber radius (af) was 950 nm, the average primary particle radius (amon) was 9.25 nm, and the filter volume fraction (Vf) was 0.050. As they are not known for the measured aggregates, we assume values of Df = 1.5, 1.75, & 2.5 with kf = 1.3, and for each assumed Df value, Nmon is calculated using Equation (8a) with the expressions provided by Thajudeen et al. (Citation2012) for Rs and PA for quasifractal aggregates: [10a]

[10b]

[10c]

[10d] [10e] [10f]

Equations (10a–f) are regression equations based upon direct calculation of both Rs and PA for computationally generated aggregates with kf = 1.3 and Df = 1.3–2.6. Subsequently, we calculate L via Equations (7a–b). is a plot of the single-fiber efficiency as a function of mobility-equivalent diameter for the unsintered aggregates (the room temperature sample) of the study by Kim et al. (Citation2009). Experimental results are plotted along with calculations for three different fractal dimensions. In addition, results are provided directly in , which additionally lists the calculated χf, Nmon, KnD, and R values for each assumed fractal dimension. Finally, in , the regression curve Ef = 0.84Pe−0.43 (where Pe = 2χf1/2/[(1+R)KnD]) is plotted, which was reported by Wang et al. (Citation2007) to agree well with experimental single-fiber efficiency measurements for the same filters used by Kim et al. (Citation2009) when challenged with spherical particles. This regression curve leads to prediction distinct from the traditional equations of depth-filtration theory as well as the equations derived from MFPT calculation. Evident in , its predictions agree well with measurements at dm = 30 and 50 nm, but do not agree with measurements of larger aggregates. Single-fiber efficiency curves predicted using MFPT-derived equations for all aggregates, conversely, are in much better agreement with measurements than the filter-specific regression equation (as well as the calculations of Kim et al. Citation[2009]). In particular, the single-fiber efficiency calculations with Df = 1.50 are within 15% of all measurements in the dm = 100–300 range. displays a calculated Nmon as a function of the mobility equivalent diameter in the dm = 100–300 range. Nmon versus dm plots yield the mass-mobility exponent, and for Df = 1.5, the inferred value of Dmm = 2.05 is in remarkably good agreement with the value of Dmm = 2.06 determined experimentally. As the single-fiber efficiency measurements and mass-mobility exponent measurements are independent of one another, we find excellent agreement with calculations for Df = 1.5, kf = 1.3 (with no other fit parameters) again supports the use of the presented dimensionless collision rate and single-fiber efficiency equations for non-spherical particles.

Table 4. A summary of the comparison between model predictions and the measurements of Kim et al. (Citation2009) for the single-fiber efficiency for a fibrous filter collecting soot aggregates. The comparison is for unsintered particles only. dm denotes the mobility equivalent diameter of the particles examined. All other dimensionless parameters are based upon the filter properties and penetration data provided by Kim et al. (Citation2009). The number of monomers per aggregate (Nmon), R, KnD, Ef (pred.), and % Diff are calculated for different assumed fractal dimensions of 1.5, 1.75, and 2.0

FIG. 6. (a) A plot of the single-fiber efficiency for soot aggregates from the experiments of Kim et al. (Citation2009) as well as predicted single-fiber efficiencies for fractal dimensions of 1.5, 1.75, and 2.5. Moreover, the empirical fit of Wang et al. (Citation2007) for the same filters with spherical particles is shown here, with which the single-fiber efficiency is calculated via the equation: Ef = 0.84Pe−0.43. (b) A plot of the number of monomers per aggregate determined from Equation (8a) as a function of mobility diameter. The resulting mass-mobility exponents (Dmm) are labeled.

FIG. 6. (a) A plot of the single-fiber efficiency for soot aggregates from the experiments of Kim et al. (Citation2009) as well as predicted single-fiber efficiencies for fractal dimensions of 1.5, 1.75, and 2.5. Moreover, the empirical fit of Wang et al. (Citation2007) for the same filters with spherical particles is shown here, with which the single-fiber efficiency is calculated via the equation: Ef = 0.84Pe−0.43. (b) A plot of the number of monomers per aggregate determined from Equation (8a) as a function of mobility diameter. The resulting mass-mobility exponents (Dmm) are labeled.

CONCLUSIONS

Following a procedure analogous to that of Thajudeen et al (Citation2012), we have shown how to apply the MFPT-inferred particle-fiber dimensionless collision kernel expression to particles of arbitrary shape, and provide algebraic equations specifically for calculations with quasifractal aggregates and cylinders. The resulting collision rate calculations are directly linked to the single-fiber efficiency. Unique from prior approaches wherein non-spherical particle filtration was examined (Balazy and Podgorski Citation2007), in the presented approach, particle geometries were taken into account explicitly, and the effective collision length scales for particles were calculated directly (not assumed). This leads to different single-fiber efficiency equations than have been developed elsewhere. Predicted single-fiber efficiencies are in reasonable agreement with those measured for well-characterized carbon nanotubes and good agreement with those measured for soot aggregates with a fit fractal dimension. The equations developed here should aid in analyzing measurements of non-spherical particle penetration through fibrous filters. Future applications of the MFPT approach should additionally enable determination of the collision rate between particles and non-spherical filter fibers, or with non-fibrous media (e.g., sintered metal spheres).

ACKNOWLEDGMENTS

The authors thank the University of Minnesota Supercomputing Institute (MSI) for providing the high performance computing resources used in MFPT calculations.

Additional information

Funding

This work was supported by the University of Minnesota Center for Filtration Research (CFR).

REFERENCES

  • Balazy, A., and Podgorski, A. (2007). Deposition Efficiency of Fractal-like Aggregates in Fibrous Filters Calculated using Brownian Dynamics Method. J. Colloid Interf. Sci. 311:323–337.
  • Chakrabarty, R. K., Moosmuller, H., Arnott, W. P., Garro, M. A., Tian, G. X., Slowik, J. G., et al. (2009). Low Fractal Dimension Cluster-Dilute Soot Aggregates from a Premixed Flame. Phys. Rev. Lett. 102.
  • Dahneke, B. E. (1973). Slip Correction Factors for Nonspherical Bodies- III The Form of the General Law. J. Aerosol Sci. 4:163–170.
  • Dahneke, B. E. (1983). Simple Kinetic Theory of Brownian Diffusion in Vapors and Aerosols, in Theory of Dispersed Multiphase Flow, R. E. Meyer, ed., Academic Press, New York.
  • Douglas, J. F., Zhou, H. X., and Hubbard, J. B. (1994). Hydrodynamic Friction and the Capacitance of Arbitrarily-Shaped Objects. Phys. Rev. E 49:5319–5337.
  • Filippov, A. V., Zurita, M., and Rosner, D. E. (2000). Fractal-like Aggregates: Relation Between Morphology and Physical Properties. J. Colloid Interf. Sci. 229:261–273.
  • Gopalakrishnan, R., and Hogan, C. J. (2011). Determination of the Transition Regime Collision Kernel from Mean First Passage Times. Aerosol Sci. Technol. 45:1499–1509.
  • Gopalakrishnan, R., Meredith, M. J., Larriba-Andaluz, C., and Hogan, C. J. (2013a). Brownian Dynamics Determination of the Bipolar Steady State Charge Distribution on Spheres and Non-spheres in the Transition Regime. J. Aerosol Sci. 63:126–145.
  • Gopalakrishnan, R., Thajudeen, T., and Hogan, C. J. (2011). Collision Limited Reaction Rates for Arbitrarily Shaped Particles Across the Entire Diffusive Knudsen Number Range. J. Chem. Phys. 135:054302.
  • Gopalakrishnan, R., Thajudeen, T., Ouyang, H., and Hogan Jr, C. J. (2013b). The Unipolar Diffusion Charging of Arbitrary Shaped Aerosol Particles. J. Aerosol Sci. 64:60–80.
  • Hansen, S. (2004). Translational Friction Coefficients for Cylinders of Arbitrary Axial Ratios Estimated by Monte Carlo Simulation. J. Chem. Phys. 121:9111–9115.
  • Huang, H., Oh, C., and Sorensen, C. M. (1998). Structure Factor Scaling in Aggregating Systems. Phys. Rev. E 57:875–880.
  • Hunt, B., Thajudeen, T., and Hogan, C. J. (2014). The Single-Fiber Collision Rate & Filtration Efficiency for Nanoparticles I. The First Passage Time Calculation Approach. Aerosol Sci. Technol. In Press.
  • Kasper, G., and Shaw, D. T. (1982). Comparative Size Distribution Measurements on Chain Aggregates. Aerosol Sci. Technol. 2:369–381.
  • Kim, S. C., Wang, J., Emery, M. S., Shin, W. G., Mulholland, G. W., and Pui, D. Y. H. (2009). Structural Property Effect of Nanoparticle Agglomerates on Particle Penetration through Fibrous Filter. Aerosol Sci. Tech. 43:344–355.
  • Kirsch, A. A., Stechkina, I. B., and Fuchs, N. A. (1974). Gas Flow in Aerosol Filters Made of Polydisperse Ultrafine Fibres. J. Aerosol Sci. 5:39–45.
  • Kuwabara, S. (1959). The Forces Experienced by Randomly Distributed Parallel Circular Cylinders or Spheres in a Viscous Flow at Small Reynolds Numbers. J. Phys. Soc. Japan 14:527–532.
  • Latin, D., Ferry, D., Gay, J. M., Delhaye, D., and Ouf, F. X. (2013). On methods Determining the Fractal Dimension of Combustion Aerosols and Particle Clusters. J. Aerosol Sci. 58:41–49.
  • Lee, K. W., and Liu, B. Y. H. (1982). Theoretical-Study of Aerosol Filtration by Fibrous Filters. Aerosol Sci. Technol. 1:147–161.
  • Narsimhan, G., and Ruckenstein, E. (1985). Monte-Carlo Simulation of Brownian Coagulation over the Entire Range of Particle Sizes from Near Molecular to Colloidal - Connection Between Collision Efficiency and Interparticle Forces. J. Colloid Interf. Sci. 107:174–193.
  • Scheckman, J. H., McMurry, P. H., and Pratsinis, S. E. (2009). Rapid Characterization of Agglomerate Aerosols by In Situ Mass-Mobility Measurements. Langmuir 25:8248–8254.
  • Schmidt-Ott, A., Baltensperger, U., Gaggeler, H. W., and Jost, D. T. (1990). Scaling Behaviour of Physical Parameters Describing Agglomerates. J. Aerosol Sci. 21:711–717.
  • Seto, T., Furukawa, T., Otani, Y., Uchida, K., and Endo, S. (2010). Filtration of Multi-Walled Carbon Nanotube Aerosol by Fibrous Filters. Aerosol Sci. Technol. 44:734–740.
  • Sorensen, C. M. (2011). The Mobility of Fractal Aggregates: A Review. Aerosol Sci. Technol. 45:765–779.
  • Sorensen, C. M., Kim, W., Fry, D., Shi, D., and Chakrabarti, A. (2003). Observation of Soot Superaggregates with a Fractal Dimension of 2.6 in Laminar Acetylene/Air Diffusion Flames. Langmuir 19:7560–7563.
  • Stechkina, I. B., and Fuchs, N. A. (1966). Studies on Fibrous Aerosol Filters—I. Calculation of Diffusional Deposition of Aerosols in Fibrous Filters. Ann. Occup. Hyg. 9:59–64.
  • Tajima, N., Fukushima, N., Ehara, K., and Sakurai, H. (2011). Mass Range and Optimized Operation of the Aerosol Particle Mass Analyzer. Aerosol Sci. Technol. 45:196–214.
  • Thajudeen, T., Gopalakrishnan, R., and Hogan, C. J. (2012). The Collision Rate of Nonspherical Particles and Aggregates for all Diffusive Knudsen Numbers. Aerosol Sci. Technol. 46:1174–1186.
  • Thajudeen, T., and Hogan, C. J. (2012). First Passage Calculation of the Conductivity of Particle Aggregate-Laden Suspensions and Composites Powder Technol. 218:31–39.
  • Wang, J., Chen, D. R., and Pui, D. Y. H. (2007). Modeling of Filtration Efficiency of Nanoparticles in Standard Filter Media. J. Nanoparticle Res. 9:109–115.
  • Wang, J., Kim, S. C., and Pui, D. Y. H. (2011a). Carbon Nanotube Penetration through a Screen Filter: Numerical Modeling and Comparison with Experiments. Aerosol Sci. Technol. 45:443–452.
  • Wang, J., Kim, S. C., and Pui, D. Y. H. (2011b). Measurement of Multi-wall Carbon Nanotube Penetration Through a Screen Filter and Single-fiber Analysis. J. Nanoparticle Res. 13:4565–4573.
  • Zhang, C., Thajudeen, T., Larriba, C., Schwartzentruber, T. E., and Hogan, C. J. (2012). Determination of the Scalar Friction Factor for Non-spherical Particles and Aggregates Across the Entire Knudsen Number Range by Direct Simulation Monte Carlo (DSMC). Aerosol Sci. Technol. 46:1065–1078.
  • Žukauskas, A. (1972). Heat Transfer from Tubes in Crossflow. Adv. in Heat Transfer 8:93–160.
  • Zurita-Gotor, M., and Rosner, D. E. (2002). Effective Diameters for Collisions of Fractal-like Aggregates: Recommendations for Improved Aerosol Coagulation Frequency Predictions. J. Colloid Interf. Sci. 255:10–26.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.