Sie sind auf Seite 1von 12

Journal of Aerosol Science 52 (2012) 45–56

Contents lists available at SciVerse ScienceDirect

Journal of Aerosol Science


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

Performance study of a disk-to-disk thermal precipitator


Bin Wang a,b, Qisheng Ou b, Shu Tao a, Da-Ren Chen b,n
a
College of Urban & Environmental Science, Peking University, Beijing 100871, China
b
Department of Energy, Environmental and Chemical Engineering, Washington University in St. Louis, Campus Box 1180,
One Brookings Drive, St. Louis, MO 63130, USA

a r t i c l e i n f o abstract

Article history: In this study, the performance of a thermal precipitator of the disk-to-disk type was
Received 23 November 2011 investigated experimentally and numerically. The prototype precipitator was basically two
Received in revised form disks separated via a circular Teflons spacer. The temperatures of the two disks (one at
28 March 2012
elevated temperature and the other at room temperature) were individually controlled by a
Accepted 3 April 2012
Available online 26 April 2012
silicone heating element and running water at room temperature. Monodisperse particles of
sodium chloride and fluorescein sodium were used to investigate the particle collection
Keywords: efficiency of the precipitator when operated under various aerosol flowrates and tempera-
Disk-to-disk thermal precipitator ture gradients. Our experimental data showed that the particle collection efficiency of the
Thermophoresis
precipitator remained approximately constant for test particles with diameters smaller than
Ultrafine particle sampler
300 nm and noticeably decreased as the particle diameter increased beyond 300 nm. A
Particle collection efficiency
numerical model was developed and showed that the calculated particle collection
efficiency was in reasonable agreement with experiment observations. Finally, a simple
model was developed to estimate the particle collection efficiency of a typical disk-to-disk
thermal precipitator. The model indicated that the particle collection efficiency of a disk-to-
disk precipitator is a function of cold-disk deposition area, the average thermophoretic
velocity, and the aerosol flowrate. This model may be useful in the future design of a
thermal precipitator with the similar configurations.
& 2012 Elsevier Ltd. All rights reserved.

1. Introduction

Much effort in ambient air pollution studies focuses on particulate matters (PMs), especially fine and ultrafine particulates
(FPs/UFPs). PMs play a critical role in global climate change (Katsouyanni et al., 1997) and human health (Dockery et al., 1993;
Pope and Dockery, 2006). Because of their ability to translocate through the epithelium of terminal bronchioles and alveoli in
the human lung system (Chalupa et al., 2004; Geiser et al., 2005), FPs/UFPs may be particularly relevant to pulmonary and
cardiovascular diseases, cancer, and mortality, as indicated by studies on both acute and long term effects (Dockery et al.,
1993; Pope and Dockery, 2006; Bräuner et al., 2009). Epidemiological studies have also evidenced that increased asthma
prevalence, including the number of patients diagnosed with the disease as well as asthma-related hospital visits, is closely
associated with UFP levels in ambient air, regional motor vehicle traffic density, and residential proximity to freeways
(Holguin, 2008; Salam et al., 2008; Patel and Miller, 2009). It is therefore important to collect fine and ultrafine aerosols and
characterize their composition, structure, and surface morphology to understand their toxicity mechanisms.

n
Corresponding author. Tel.: þ1 314 935 7924; fax: þ1 314 935 7211.
E-mail address: chen@seas.wustl.edu (D.-R. Chen).

0021-8502/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jaerosci.2012.04.004
46 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

To perform this characterization task by offline instruments such as SEM/TEM and other chemical analytic instru-
ments requires collecting ambient particles while keeping them unaltered during the sampling period (Chung et al., 2001;
Kim et al., 2001; Smith et al., 2005). Filtration and impaction techniques are the most commonly used for sampling
ambient particles. The major concern of using such techniques is the potential alternation of the morphology and structure
of agglomerated particles. Another technique for the collection of fine and ultrafine particles is electrostatic precipitation.
However, the performance of electrostatic samplers is greatly influenced by the charge state of sampled particles because
they require to be electrically charged prior to the electrostatic collection. The challenge is to control the electrical charges
on sampled particles to a level that permit the ‘‘soft’’ collection of agglomerate particles.
Sampling devices making use of thermophoresis effect, i.e., thermal precipitators, are excellent candidates for carrying
the soft collection of particles, especially agglomerates. Thermophoresis is the physical phenomenon in which particles
move from higher to lower temperature zone (i.e., in a temperature gradient field) because of the non-zero net molecular
momentum transfer on individual particles (Lin and Tsai, 2003; Tsai et al., 2004; Lin et al., 2004). The formula offered in
the work of Talbot et al. (1980) is popularly used in the literature for the calculation of particle thermophoretic velocity in
a temperature gradient field. Good agreement between experimental data and theoretical prediction on the thermopheric
velocity of particles has been reported among the literatures (Montassier et al., 1991; Tsai and Lu, 1995; Messerer et al.,
2003; Azong-Wara et al., 2009). The particle thermopheric effect has been applied both to collect PMs on substrates of
different types (Kethley et al., 1952; Orr and Martin, 1958; Montassier et al., 1991; Lorenzo et al., 2007) and to reduce
particles from the deposition to surfaces (Dedrick et al., 2005).
Thermal precipitators in various geometrical configurations have been developed over the years. The configuration
types of thermal precipitators include the hot-wire (Green and Watson, 1935), tube-to-tube (Bredl and Grieve, 1951),
disk-to-disk (Kethley et al., 1952; Orr and Martin, 1958; Wright, 1953), plate-to-plate (Tsai and Lu, 1995; Azong-Wara
et al., 2009). Among them thermal precipitators of the plate-to-plate configuration have been most widely developed,
and their performance has been intensively investigated. However, very little attention has been paid to disk-to-disk
thermal precipitators. Previous work on disk-to-disk precipitators can be found in the papers of Kethley et al. (1952), Orr
and Martin (1958) and Wright (1953). A disk-to-disk thermal precipitator essentially consists of two co-axially aligned
disks, one aerosol inlet, and one outlet (often located at the disk edge). The temperature gradient in a disk-to-disk thermal
precipitator is established by keeping two disks at different temperature settings.
Compared to plate-to-plate thermal precipitators disk-to-disk precipitators have several unique advantages: (1) the aerosol
sampling flow in the disk-to-disk precipitators can be in perfectly axisymmetric condition while the flow transition between
the round inlet/outlet and the precipitation chamber in the rectangular cross-section is necessary in plate-to-plate devices;
(2) omni-directional aerosol sampling is possible for disk-to-disk prcipiators when the aerosol sampling flow is designed to move
radially towards to the center; (3) the design concept of disk-to-disk precipitators makes it easy to integrate with cell culture
plates having multiple wells. Our ultimate goal for this investigation is to eventually integrate the design of disk-to-disk
precipitator with cell culture plates and use the integrated exposure system to study the toxicity of ambient particles via in-vitro
experiments.
Unfortunately, parametric studies on the performance of a disk-to-disk precipitator (i.e., the particle collection
efficiency as a function of temperature gradient, aerosol flowrate, and particle size) had not been carried out in
the previous studies, resulting in no design guideline for disk-to-disk thermal precipitators. To gain knowledge for
the future design of similar samplers, a disk-to-disk thermal precipitator was thus designed and fabricated in this study.
The performance of the prototype (i.e., the particle collection efficiency) was characterized by DMA-classified sodium
chloride (NaCl) and fluorescein sodium (F-Na) particles under various aerosol flowrates and temperature gradients.
Numerical modeling was also performed to predict the collection efficiency of the prototype. Additionally, a simple model
was further developed to estimate the collection efficiency of a generic disk-to-disk type thermal precipitator for the
future designs.

2. Precipitator design

A schematic diagram of our disk-to-disk thermal precipitator is shown in Fig. 1. It consists of two aluminum disks
(12.7 cm in diameter). Note that the upper (heated) and lower (cooled) disks were both 3.175 mm in thickness and the
lower disk was machined to have a recess of 1.0 mm depth in its center. At the perimeter, the two disks were separated by
a 0.35 mm thick Teflons gasket. Thus, the total gap between the inner disk surfaces was 1.35 mm. The temperatures of the
hot and cold disks were individually controlled by a silicon heating disk (Kaptons Heaters) connecting a temperature
controller (Cole-Parmer, Model: 89810-00) and running water. Aerosol flow entered the precipitator from an inlet tube
located at the upper disk’s center. The aerosol stream flowed outwards in the radial direction. Uniformly distributed near
the outer edge of the lower disk, 72 orifices (0.794 mm in diameter) introduced the aerosol flow into the annular chamber
located below the cold disk. The aerosol flow was exhausted from four outlets 901 apart (as shown in Fig. 1(b)). For each
disk, two thermocouples measured the temperatures at four locations equally spaced in the radius of the particle
precipitation area and separated at 901 apart. The 901 separation is because the size of thermocouple probes physically
limited the possibility of placing four probes in a line. In our preliminary temperature measurement we found that
monitoring the temperature at two specific locations for each disk was sufficient to ensure the temperature uniformity of
the entire disk.
B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56 47

Fig. 1. Schematic diagram of the studied disk-to-disk type thermal precipitator (not to the scale); (a) side view and (b) top view.

Electrostatic Classifier
(TSI 3080)
Laminar
flowmeter #1 Neutralizer
Neutralizer HEPA
Po210 #2 Po210 #3 Filter #2

Diffusion HEPA Filter


Valve #2
Dryer
Precipitator
Valve #1 UCPC
(TSI 3025A)
Neutralizer
Po210 #1
HEPA
Valve #3 Filter #3
Atomizer Compressed air
Vacuum
Laminar
flowmeter #2

Fig. 2. Schematic diagram of the experimental setup.

3. Experimental setup and procedure

3.1. Experiment setup

The experimental setup for calibration of prototype under various operational conditions is shown in Fig. 2. Test
particles were first generated by a custom-made Collison atomizer, and then passed first through a Po210 radioactive
neutralizer (#1) to minimize electrical charges on particles and then through a silica gel diffusion dryer (#1) to remove
solvent from the particles. To obtain monodisperse test particles, the generated particles were fed into a second Po210
neutralizer and a differential mobility analyzer (DMA) to classify particles of the desired sizes. The aerosol flowrate
entering the classifier was monitored by a calibrated laminar flowmeter and controlled by a needle valve (#1) in the
exhaust branch line upstream of the flowmeter. A HEPA cartridge filter was installed upstream of the needle valve to
prevent particles from accumulating in the valve. DMA-classified particles are singly charged. Charges on classified
particles were further reduced in a third Po210 neutralizer. The particle number concentrations both the upstream and
downstream of the prototype were measured by an ultrafine condensation particle counter (UCPC, TSI model 3025) via the
use of a three-way valve (#2). The flowrate of aerosol stream passing the prototype was varied by adding a vacuum line
(with a laminar flowmeter and needle valve) downstream. The total aerosol flowrate was the sum of the CPC sampling
flowrate and vacuum line flowrate. The CPC in our study was operated at low flow mode (0.3 lpm).

3.2. Experimental procedure

3.2.1. Collection efficiency


Sodium chloride (NaCl) and fluorescein sodium (F-Na) particles in the diameter ranging from 40 to 500 nm were used
to investigate the collection efficiency of the prototype. A differential mobility analyzer (DMA, TSI model 3081) was used
48 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

to classify particles in the above size range. For each particle size used (40, 60, 80, 100, 150, 200, 300, 400, and 500 nm), the
ratio of the DMA sheath flowrate to the volumetric aerosol flowrate was maintained at 6:1 or higher. The aerosol flowrate
though the studied precipitator was controlled via the UCPC sampling flowrate by a vacuum line downstream. The same
UCPC was used to measure the particle concentrations at upstream and downstream of the precipitator, noted as Cup and
Cdown. The particle loss in the transport tubing (Ltube) was measured by removing the precipitator from the experimental
setup, measuring the particle concentration at the same downstream and upstream sampling ports, and then calculating
the ratio of the measured concentrations. The total particle loss (or collection) in the disk-to-disk thermal precipitator
(noted as L) was further derived as:
C down
L ¼ 1 ð1Þ
C up  ð1Ltube Þ

The particle loss in the prototype due to the particle Brownian diffusion and gravitational settling in the precipitator at
various temperature gradients could not be directly measured. For reference, we measured the above loss (noted as Ldiff) at
three different aerosol flowrates under no temperature gradient. The collection efficiency due to particle thermophoresis
(noted as Zth) was then calculated by L–Ldiff, where L is the total loss of particles in the thermal precipitator under a specific
temperature gradient.

3.2.2. Deposition distribution


To examine the spatial distribution of particles deposited on the cold disk surface, a layer of thin aluminum foil
(thickness ¼0.02 mm) was glued on the cold disk via PTFE grease (Saunders Enterprises, Inc.). DMA-classified F-Na
particles with diameters of 200 nm were used in this part of the experiment. The experimental setup shown in Fig. 1 run
for approximately 8 h. The aluminum foil was removed after the deposition and cut into one circular and four annular
pieces using 4 punches with the diameters of 1.27, 2.54, 4.76, and 8.26 cm. The five aluminum pieces were individually
immersed in five 50-ml beakers, each containing 20 ml ultrapure water, and sonicated for 30 s. The inner surfaces of both
inlet and outlet tubes of the prototype were also washed by 20 ml of ultrapure water. The fluorescent intensity of each
dissolved solution was then measured by a Turner Biosystems Modulus Fluoro Lumino Photomtr 9200-003 (Turner
Biosystems Inc.), and their values were noted as F1, F2, F3, F4, and F5 (for the five aluminum pieces, respectively). The
percentage of particle mass deposited (Pi) on each designated area of the cold disk was then calculated as
F
P i ¼ P5 i for i ¼ 125 ð2Þ
j¼1 Fj

4. Numerical modeling

A numerical model was also developed to predict the particle collection efficiency of this prototype. In this model the
flow field and temperature distribution in the precipitation zone were modeled via the COMSOL finite element package
(Version 4.0a, COMSOL, Inc.). The particle trajectory calculation was then implemented by solving the particle Lagenvin
equation. The overall collection efficiency could thus be derived from the fate of particles entering the collection zone of
the prototype. The computational domain and a typical mesh distribution in the domain are shown in Fig. 3, and the
COMSOL boundary modules and boundary conditions were described in Table 1. Because of axisymmetrical configuration
of the prototype we used only half of the precipitation zone as the computational domain. A total of approximate 25,000
elements were used in the modeling.
In this model the flow field in the precipitator was assumed to be steady-state, incompressible, and axisymmetric. The
incompressible Navier–Stokes and the continuity equations were thus used to model the flow field. Lagrangian elements
with polynomials of the 2nd order for velocity and the 1st order for pressure were used for the finite-element computation
of flow field. A uniform airflow velocity profile was applied at the inlet of the computational domain, and the pressure at a
point on the boundary was specified as zero for the reference. The no-slip boundary condition was imposed on the wall
boundary. Zero gradients in the airflow velocity and pressure were applied along the center line of the computational
domain. The temperature distribution in the precipitator was modeled by solving the heat convective-diffusion equation.
Based on the experimental measurement specific temperatures were imposed on the hot and cold disks of the
computational domain. The incoming flow was assumed to be at the room temperature, and no temperature profile
was specified at the outlet. A zero gradient in the temperature was also applied at the centerline of the domain.
Once the flow field and temperature fields were calculated, particles were assumed to be uniformly released from the
entrance of inlet tube, and the trajectory of each particle was calculated by the following equation:
,
@u , , , ,
mp ¼ F external ¼ F th þF drag þ G ð3Þ
@t
,
where mp and up
are particle mass and airflow velocity, respectively. Three external forces acting on particles were taken
into the consideration in this model: particle drag force, the thermophoretic effect and gravitational settling. With the
B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56 49

Fig. 3. (a) The computational domain and (b) a typical mesh distribution used in the domain for the numerical modeling of the disk-to-disk thermal
precipitator.

Table 1
Physical modules and boundary conditions used in the numerical modeling of disk-to-disk thermal precipitator.

Physical module Boundary condition Boundaries

Laminar flow Axial symmetric AI


No slip wall BC, CD, DE, EF, GH, HI
Uniform velocity inlet AB
Outlet with pressure constrait at one point FG

Heat transfer Axial symmetric AI


Thermal insulation DE
Constant temperature AB, BC, EF, FG, HG
Measured temperature profile CD, HI

assumption of spherical particles, the drag force acting on an individual particle can be estimated as

prg C D d2 V 2
F drag ¼ ð4Þ
8C C

where rg is the air density, d is the particle diameter, V is the relative velocity between air and particle, and CD is the drag
coefficient, which can be expressed as
24
CD ¼ ð1 þ 0:15Re0:687 Þ ð5Þ
Re
where Re is the particle Reynolds number. CC is the Cunningham correction factor (Hinds, 1999), and can be estimated as
  
l l
CC ¼ 1 þ 2:34 þ1:05exp 0:39 ð6Þ
d d

where l is the mean free path of air at ambient temperature and pressure.
The thermophoretic force acting on a particle with a diameter dp under a temperature gradient rT can be described by
the following expression (Batchelor and Shen, 1985):
3pm2 dp H
F th ¼ rT ð7Þ
rT
where T is the absolute temperature of the particle, and H is the thermophoretic coefficient, which depends on both air and
particle properties. An expression of H for spherical particles is given by Talbot et al. (1980) as
2C s ððkg =kp Þ þ C t KnÞC C
H¼ ð8Þ
ð1 þ 3C m KnÞð1þ ð2kg =kp Þ þ2C s KnÞ
50 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

where Cs ¼1.147, Ct ¼ 2.20, and Cm ¼1.146. These values are the thermal creep coefficient, the thermal jump coefficient, and
the velocity jump coefficient, respectively (Batchelor and Shen, 1985), and kg and kp are the thermal conductivities of the
fluid and particle, respectively.
The Runge–Kutta method of the 4th order with variable time steps was applied to solve Eq. (3). During the trajectory
calculation, the time increment in each step was varied to ensure a particle took at least two steps to cross a single mesh element.
Note that the Brownian motion of particles was neglected in this model. It is because the characteristic diffusion length
for particles in the diameter of 40 nm (i.e, the smallest particle size tested) is less than 13% of the characteristic
thermophoretic length of the same particles in the worst case at the aerosol flowrate of 0.3 lpm and the temperature
gradient of 130 1C/cm. The above percentage is reduced to less than 5% for the cases at the aerosol flowrate of 1.5 lpm.

5. Results and discussion

5.1. Particle loss under no temperature gradient

Fig. 4 shows the particle loss under no temperature gradient in the prototype for particles of two test materials. Because
of the size range of test particles, it is reasonable to assume that particle diffusion is the dominant mechanism in particle
loss. As expected, the loss increased as the particle size decreased. For the case of 0.5 lpm aerosol flowrate, 12% particle loss
was encountered for 40-nm NaCl and F-Na particles. The loss decreased to less than 5% for particles with diameters larger
than 300 nm. The particle loss also decreased as the aerosol flowrate increased.

5.2. Collection efficiency under temperature gradients

5.2.1. Material dependence


Fig. 5 shows the collection efficiency of both NaCl and F-Na particles with diameters ranging from 40 to 500 nm at three
aerosol flowrates and under two temperature gradients. For the case of 0.5 lpm aerosol flowrate and a 248 1C/cm
temperature gradient, no significant difference was observed in the collection efficiency between two particle materials.
However, an efficiency difference was observed for the same flowrate and under the 130 1C/cm gradient. For F-Na particles
with diameters of 400 and 500 nm, they had lower collection efficiency than NaCl particles with the same diameters. Less
than an 8% difference in the particle collection efficiency between particles of the two materials was observed for
aerosol flowrates of 1.0 and 1.5 lpm under the two tested temperature gradients. Similar observation was
also found in a plate-to-plate thermal precipitator (Tsai and Lu, 1995). The thermal conductivities of the NaCl (Reist, 1993)
and F-Na (Montassier et al., 1991) particles were about 6.5 and 0.43 W/(m K), respectively. The significant difference in the
thermal conductivity of two test particle materials was supposed to result in the observable difference in the thermophoretic
velocity of particles at the same sizes but made of two different materials. However, the above deduction is based on the
theory for particles in the continuum regime (i.e., small particle Knudsen number). As the particle size is reduced (i.e.,
increasing the particle Knudsen number) the momentum and heat transfers surrounding particles are in the transitional
regime (in which the velocity and heat flux jumps at the particle surface are present). As a result the particle material effect
becomes less significant for small particles. Since all test particles were less than 500 nm in diameter the effect of material on
the particle thermophoretic velocity is much reduced (Brock, 1962; Talbot et al., 1980; Santachiara et al., 2002). Indeed the
difference in the themophoretic velocities of NaCl and F-Na particles calculated by the formula given by Talbot et al. (1980)
were found to be less than 5% for diameters ranging from 40 to 500 nm. It is the reason why the particle collection efficiency
of the studied precipitator is marginally dependent on the particle material.

15% 15%
Particle losss percent
Particle loss percent

10% 10%
0.5 lpm
1.0 lpm
5% 5% 1.5 lpm

0% 0%
10 100 1000 10 100 1000
Diameter/nm Diameter/nm

Fig. 4. The particle loss percent of NaCl particles (A) and F-Na particles (B) (40 nmrdiameter r500 nm) at 3 inlet air flowrates under no temperature
difference in the precipitator. The space between the two disks was 1.35 mm.
B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56 51

100% 100%

80%
Collection efficiency

Collection efficiency
80%
248 °C/cm (NaCl) 248 °C/cm (NaCl)
60% 60%
248 °C/cm (F-Na) 248 °C/cm (F-Na)
40% 130 °C/cm (NaCl) 40% 130 °C/cm (NaCl)
130 °C/cm (F-Na) 130 °C/cm (F-Na)
20% 20%

0% 0%
10 100 1000 10 100 1000
Diameter/nm Diameter/nm

100%
Collection efficiency

80%
248°C/cm (NaCl)
60%
248°C/cm (F-Na)
40% 130°C/cm (NaCl)
130°C/cm (F-Na)
20%

0%
10 100 1000
Diameter/nm

Fig. 5. The collection efficiency of NaCl and F-Na particles (40 nm rdiameter r500 nm) under two temperature gradients at 3 inlet air flowrates,
(A) (0.5 lpm), (B) (1.0 lpm), and (C) (1.5 lpm). The space between the two disks was 1.35 mm.

5.2.2. Effect of particle size


The collection efficiency of the prototype gradually decreased with an increase in particle size. The collection efficiency
for both NaCl and F-Na particles deceased by 15–25% over the particle size range from 40 to 500 nm when the prototype
was operated at three aerosol flowrates (i.e., 0.5, 1.0 and 1.5 lpm) and under 130 and 248 1C/cm temperature gradients.
Under the 248 1C/cm gradient, the maximal particle collection efficiency of the prototype operated at the 0.5 lpm aerosol
flowrate was approximately 100% for particles less than 150 nm in size. Under both tested temperature gradients, the
particle collection efficiency at the 0.5 lpm flowrate was then dramatically decreased for particles with diameters larger
than 300 nm. A similar observation was found in the work of Tsai and Lu, (1995), in which a plate-to-plate precipitator and
DMA-classified particles were used. However, the observed size-dependence on the collection efficiency was not obvious
in the work of Messerer et al. (2003), in which polydisperse soot particles and a plate-to-plate type precipitator were used.
Possibly the use of polydisperse particle masked the size dependence effect, and the fact that the 1555 1C/cm gradient used
in the Messerer’s work was sufficiently high to collect all the particles, resulted in less chance to observe the trend. It is
also probable that the thermal properties of soot particles differ greatly from those of the particles tested in our work.

5.2.3. Effect of aerosol flowrate


As shown in Fig. 5, the measured particle collection efficiency of the prototype decreased as the aerosol flowrate
increased, under both tested temperature gradients for particles of both materials. Further, the decrease in the collection
efficiency from the case of the 0.5 lpm flowrate to that of 1.0 lpm is more obvious than that from 1.0 to 1.5 lpm. It could be
because the particle residence time in the precipitation zone was decreased by a factor of two in the former case and the
time merely decreased by 50% in the later case. We excluded the effect of pressure drop for the aforementioned
observation because the pressure drop of the studied precipitator at the 1.5 lpm airflow (i.e., the maximal flowrate used in
our study) was measured at slightly less than 0.2 in H2O, which is negligible when compared with the absolute ambient
pressure. We also found that the collection efficiency of the prototype disk-to-disk precipitator decreased more gradually
than that of a plate-to-plate precipitator for both Fe and NaCl particles (Gonzalez et al., 2005).

5.3. Particle deposition distribution on cold disk

Fluorescence analysis was also applied to examine the deposition uniformity of particles deposited on the cold disk
surface. In this part of the study, F-Na particles with the diameter of 200 nm were used and the prototype was operated at
the aerosol flowrate of 0.5 lpm and the two temperature gradients. The measured percentages of particle mass deposited
in all designated areas of cold disk is shown in Fig. 6. The numerical values of both the percentage of particle mass
deposited in each area and the deposition percentage per unit area are also given in Table 2 for the reference. It is evident
that the percentage of particle mass collected on the cold disk under the two tested temperature gradients in the spacing
between two disk has similar trends (p 40.05). More deposition percent per unit deposition area was observed in the
52 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

Fig. 6. The experimental percentage of particle mass deposited in each designated areas covering cold disk of the precipitator. F-Na particles with the
diameter of 200 nm were used. Aerosol flowrate was 0.5 lpm. The space between the two disks was 1.35 mm. Two temperature gradients in the spacing
between two disks were tested. Note that Area 1 covers the central area of cold disk (i.e., with the radius of 6.30 cm); Area 2 is the annular area of cold
disk with the radius from 0.63 to 1.27 cm; Area 3 is the annular area with the radius from 1.27 to 2.38 cm; Area 4 is that with the radius from 2.38 to
4.13 cm; and Area 5 is that with the radius from 4.13 to 6.30 cm.

Table 2
Percentages of particle mass deposited on all the designated areas which covered the cold disk in the experiment for particle deposition distribution. The
test particle size was 200 nm and the aerosol flowrate was 0.5 lpm. Two temperature gradients (i.e., 130 and 248 1C/cm) in the spacing between two disks
were tested.

Area (no.) Radius (cm) Percentage of mass deposition Percentage per unit area

130 1C/cm (%) 248 1C/cm (%) 130 1C/cm (%) 248 1C/cm (%)

1 0.00–0.63 1.90 2.00 1.52 1.60


2 0.63–1.27 3.95 4.72 1.03 1.24
3 1.27–2.38 13.39 12.53 1.05 0.99
4 2.38–4.13 38.04 38.58 1.06 1.08
5 4.13–6.30 42.72 42.17 0.24 0.24

central area with a radius up to 2.54 cm as compared with that in the area with the radius larger than 2.54 cm. The
deposition data shows no dependence on the temperature gradient at either aerosol flowrate.

5.4. Comparison with numerical prediction

Fig. 7 shows the calculated and experimental collection efficiencies of the prototype for F-Na particles at the tested
aerosol flowrates. For the high temperature gradient, good agreement between numerical and measured efficiencies for
particles with diameters less than 300 nm was obtained at the three aerosol flowrates. For the low temperature gradient
reasonable agreement in both collection efficiencies was also achieved in the same particle size range. The average
difference between the measured and calculated data was found to be less than 4% for particles r300 nm. The calculated
collection efficiency of the prototype is apparently higher than that measured for the particles Z300 nm. The maximum
discrepancy between measured and calculated efficiencies was about 10% at the particle size of 500 nm when the
prototype was operated at the 0.5 lpm aerosol flowrate and high temperature gradient. The maximal discrepancy between
the experimental and numerical efficiency at the range (300–500 nm) was generally observed in all the comparison cases
shown in Fig. 7.

6. A simple model for future design of disk-type thermal precipitators

In addition to the numerical modeling we further developed a simple model based on the operational principle of disk-
to-disk precipitators to relate the particle collection efficiency to key variables involved in the operation of disk-to-disk
precipitators.
It is assumed that the flow is steady-state, incompressible, laminar, and axisymmetric. Shown in Fig. 8 is a schematic
diagram of simplified disk-to-disk precipitator geometry in cylindrical coordinates. The thermophoretic velocity of a
particle is assumed to remain unchanged in the deposition process and be equal to that calculated using the average
temperature of the hot and cold disks. Based on Batchelor and Shen (1985), the thermophoretic velocity (Uth) can be
B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56 53

100% 100%

80% 80%

Collection efficiency
Cllection efficiency

Exp- 0.5 lpm Exp-0.5 lpm


60% Exp- 1.0 lpm 60% Exp-1.0 lpm
Exp- 1.5 lpm Exp-1.5 lpm

40% Cal-0.5 lpm 40% Cal-0.5 lpm


Cal-1.0 lpm Cal-1.0 lpm
Cal-1.5 lpm Cal-1.5 lpm
20% 20%

0% 0%
10 100 1000 10 100 1000
Diameter/nm Diameter/nm

100% 100%

80% 80%

Collection efficiency
Cllection efficiency

Exp- 0.5 lpm Exp-0.5 lpm


60% Exp- 1.0 lpm 60% Exp-1.0 lpm
Exp- 1.5 lpm Exp-1.5 lpm
40% Cal-0.5 lpm 40% Cal-0.5 lpm
Cal-1.0 lpm Cal-1.0 lpm
Cal-1.5 lpm Cal-1.5 lpm
20% 20%

0% 0%
10 100 1000 10 100 10 00
Diameter/nm Diameter/nm

Fig. 7. Comparison of numerical and experimental collection efficiencies for the cases of F-Na and NaCl particles (40 nmr diameterr 500 nm) at three
aerosol flowrates under two temperature gradients. The space between the two disks was 1.35 mm. (A) is for F-Na particles at 130 1C/cm; (B) for F-Na
particles at 248 1C/cm); (C) for NaCl particles at 130 1C/cm and (D) for NaCl particles at 248 1C/cm. Note that ‘‘Exp’’ was the experimental data and ‘‘Cal’’
the calculated data using the numerical model.

Inlet

A (Rin, z*) *

H
o
r
B (R, 0)

Fig. 8. The cylindrical coordinate system of the disk type precipitator for simple model development. The z axis is the central axis of the inlet tube, r axis
is the radius direction along the cold disk surface, H is the space between the two disks, Rin is the radius of the inlet tube, and R is the radius of the cold
disk. The particle trajectory (c*) starting from the critical height of z* and ending at the outer edge of the cold disk.

rewritten as the following:


H 0 n0
U th ¼  Ur T ð9Þ
T0
where, T0 is the average temperature of the hot and cold disks of the disk-type precipitator, and H0 and n0 are the
thermophoretic coefficient and kinematic viscosity of the air at T0. Eq. (9) can be further written as
U th ¼ Zth rT ð10Þ

where, Zth ¼ H0  n0/T0. Neglecting the inertial and Brownian motion and gravitational settling of particles during the
transport in a disk-type precipitator, the particle trajectory in the Z–R plane is then governed by:
( dr
dt
¼ ur þ Z th rT r
dz ð11aÞ
dt
¼ uz þ Z th rT z
54 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

where, ur and uz are the flow velocity components in the r- and z-directions, respectively; and rTr and rTz are the
temperature gradients in the r- and z-directions, respectively. Based on the steady heat convective-diffusion equation, one
can obtain the following governing equation for the temperature distribution (referring to Fig. 7 for the coordinate system)
ur @T @2 T
¼ 2 ð12Þ
a @r @z
where, a ¼ rgCp/kg. Integrate Eq. (12) in the range 0 rz rH, and it yields
Z      
1 H @T @T  @T 
rur dz ¼  r ð13Þ
a 0 @r @z H @z 0
It is also assumed that the airflow velocity in the z direction is close to zero because of the narrow spacing between two
disks. According to the continuity equation (Bird et al., 2002), the air flowrate in the r direction can be found as
f ðzÞ
ur ¼ ð14Þ
r
Combining Eqs. (16),(17) yields
Z    
1@ H @T  @T 
ðrur TÞ dz ¼    r ð15Þ
a @r 0 @z H @z 0
Define the mixing temperature Tm and average flow velocity at any specific r (ur ) as
RH
rur T dz
Tm ¼ 0 ð16Þ
rur H
RH
0 ur dz Q in
ur ¼ ¼ ð17Þ
H 2prH
Eq. (18) can therefore be written as the following
 
Q in @ðT m Þ @T  @T 
U ¼    ð18Þ
2par @r @z @z H 0

To further simplify Eq. (18) we define the non-dimensional parameters as


z r TT c T m T c
z¼ ; r¼ ; T¼ ; Tm ¼ ð19Þ
H R T h T c T h T c
where, Th and Tc are the temperatures of the hot and cold disks. Eq. (18) can then be expressed as
" #  
HQ in 1 @ðT m Þ @T  @T 
U ¼    ð20Þ
2paR2 r @r @z  @z 
1 0

Note that the magnitude of the coefficient (HQin/(2paR2)) in the left-hand side of Eq. (20) under the three inlet flowrates
(1.0, 1.5, and 2.0 lpm) are generally O(  1). The magnitude of the dimensionless
  convective term and the two terms in the
right-hand side of Eq. (20) are O(1). It is then concluded that @T=@z1  @T=@z0 ; i.e., the temperature distribution in the z
direction is approximately linear, and the temperature gradient in the r direction is negligible compared with that in the z
direction. With the above approximation, the particle trajectory Eq. (11a), can be calculated as
( dr
dt
¼ ur
dz ð11bÞ
dt
¼ Z th rT z

Since Brownian particle motion is not considered herein there exists a limiting particle trajectory (cn), starting from the
critical height of zn and ending at the outer edge of the cold disks (shown in Fig. 8). Particles entering the precipitation zone
from a height position less than zn will be deposited on the surface of the cold disks, and ones with an entrance position
higher than zn will escape collection. With the assumption of uniform particle concentration at the entrance, the
volumetric flowrate in which all the particles will be collected can be calculated by combining Eqs. (11b) and (14):
R zn Z zn
0 ur ðRin Þ dz
Q collect ¼ 2pRin zn ¼ 2p Rin ur ðr ¼ Rin Þ dz
zn
0
Z zn Z R
¼ 2p f ðzÞ dz ¼ 2p Z th rT z r dr ¼ pðR2 R2in ÞU zth ð21Þ
0 Rin

As the ratio of R2in =R2 is 0.5% for our prototype its collection efficiency can be approximated as
Q collect pR2 U zth
Z¼ ¼ ð22Þ
Q in Q in
B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56 55

100% 100%

80% 80%

Collection efficiency
Collection efficiency

Sim-0.5 lpm Sim-0.5 lpm


60% Sim-1.0 lpm 60% Sim-1.0 lpm
Sim-1.5 lpm Sim-1.5 lpm
40% Num-0.5 lpm 40% Num-0.5 lpm
Num-1.0 lpm Num-1.0 lpm
20% 20%
Num-1.5 lpm Num-1.5 lpm

0% 0%
10 100 1000 10 100 1000
Diameter/nm Diameter/nm

Fig. 9. Comparison of collection efficiency differences obtained by the numerical model and simple model derived for the cases of NaCl particles at two
temperature gradients (A) 130 1C/cm and (B) 248 1C/cm.The space between the two disks was 1.35 mm. Note that ‘‘Sim’’ refers ‘‘simple model and ‘‘Num’’
for numerical model.)

1.2 1.2

1.0 1.0

0.8 0.8
UthT0 / (νͪTz)
UthT0 / (νͪTz)

0.6 0.6

0.4 NaCl(130°C /cm) 0.4


F-Na l(130°C /cm)
NaCl(248°C /cm)
F-Na l(248°C /cm)
0.2 Talbot et al. (1980) 0.2 Talbot et al. (1980)
Derjaguin et al. (1976) Derjaguin et al. (1976)
0.0 0.0
0 1 2 3 4 0 1 2 3 4
Kn Kn

Fig. 10. Comparison of experimental dimensionless thermophoretic velocity UthT0nDTz with those calculated by existed theories: (a) NaCl particles
(b) F-Na particles.

Fig. 9 shows the comparison of the particle collection efficiencies calculated by Eq. (22) and the numerical modeling for
the prototype. Reasonable agreement was obtained in this comparison for particle sizes ranging from 40 to 500 nm. The
maximal difference in the efficiency between two sets of data is typically less than 2%. Note that the dimensionless
thermophoretic velocity (UthT0nDTz) can be calculated via Eq. (9) with the measured collection efficiency. The accumulated
collection efficiency along the radius of the cold calculated using Eq. (22) was also shown in Fig. 6. When the radius
o2.5 cm the calculated result was slightly higher than the experimental data. It may be caused by the airflow turbulence
in this area. When the radius 44 cm, the experiment and calculated data were of reasonable agreement.
Fig. 10 shows the above-derived thermophoretic velocity, UthT0nDTz, and that calculated by existing theories (Talbot
et al., 1980 and Derjaguim et al., 1976) for NaCl and F-Na particles. It is evident that the measured thermophoretic
velocities were closer to that given by Talbot et al. (1980) compared to that by Derjaguin et al. (1976). A similar result has
been observed in the work of Tsai and Lu, (1995) for a thermal precipitator in the plate-to-plate configuration.

7. Conclusion

In this study a disk-type thermal precipitator was developed and its performance was investigated both experimentally and
numerically. The particle collection efficiency of the prototype under two temperature gradients (130 and 248 1C/cm) at three
flowrates (0.5, 1.0 and 1.5 lpm) were measured using DMA-classified particles of sodium chloride and fluorescein sodium with
sizes ranging from 40 to 500 nm. It was found that the particle material’s effect on the collection efficiency of the prototype was
marginal. For particles with diameters r300 nm the particle collection efficiency of the prototype was nearly independent of
the particle size. However, the moderate size effect on the collection efficiency of the prototype was observed in the size range
Z300 nm (i.e., 15–25% efficiency decrease as the particle size increased from 300 to 500 nm). The fluorescence method was also
applied to examine the uniformity of particle deposited on the cold disk surface. The particle deposition on the disk surface was
uniform in the radial area from 1.0 to 2.5 in. under their two tested temperature gradients at the aerosol flowrate of 0.5 lpm.
A numerical model was developed using COMSOL to predict the performance of the prototype. The calculated collection
efficiency was in reasonable agreement with the experimental data for particles with diameters r300 nm. A noticeable
56 B. Wang et al. / Journal of Aerosol Science 52 (2012) 45–56

discrepancy between the experimental and numerical collection efficiencies was observed for particles Z300 nm. A
simple model was further developed to estimate the particle collection efficiency of a thermal precipitator of the disk type.
The simple model was validated by comparing the collection efficiency calculated by both the numerical model and the
simple one. The collection efficiency estimated by the simple model was in excellent agreement with that calculated by
the numerical model. Via the simple model, it was found that the collection efficiency was primarily determined by the
thermophretic velocity of particles, the aerosol carrier flowrate, and the surface area for particle collection.

Acknowledgment

Mr. Wang would like to express the deepest appreciation for the financial support provided by the China Scholarship Council,
which enabled him to work in the Particle Laboratory, at Washington University in St. Louis. Thanks are also offered to Mr.
Tandeep Singh Chadha, Mr. Fei Xia, Mr. Nathanael Connesson, Ms. Qiaoling Liu, and Dr. Li Huang for their thoughtful discussion
and advice about the research.

References

Azong-Wara, N., Asbach, C., Stahlmecke, B., Fissan, H., Kaminski, H., Plitzko, S, & Kuhlbusch, T.A.J. (2009). Optimisation of a thermophoretic personal
sampler for nanoparticle exposure studies. Journal of Nanoparticle Research, 11, 1611–1624.
Batchelor, G.K., & Shen, C. (1985). Thermophoretic deposition of particles in gas flowing over cold surfaces. Journal of Colloid and Interface Science, 107, 21–37.
Bird, R.B., Stewart, W.E., & Lightfoot, E.N. (2002). Transport Phenomena (2nd ed.). John Wiley & Sons, Inc.: New York, pp. 846.
Bräuner, E.V., Mortensen, J., Møller, P., Bernard, A., Vinzents, P., Wåhlin, P., Glasius, M., & Loft, S. (2009). Effects of ambient air particulate exposure on
blood-gas barrier permeability and lung function. Inhalation Toxicology, 21, 38–47.
Bredl, J., & Grieve, T.W. (1951). A thermal precipitator for the gravimetric estimation of solid particles in flue gases. Journal of Scientific Instruments, 28, 21–23.
Brock, J.R. (1962). On the theory of thermal forces acting on aerosol particles. Journal of Colloid Science, 17, 768–780.
Chalupa, D.C., Morrow, P.E., Oberdörster, G., Utell, M.J., & Frampton, M.W. (2004). Ultrafine particle deposition in subjects with asthma. Environmental
Health Perspectives, 112, 879–882.
Chung, A., Chang, D.P., Kleeman, M.J., Perry, K.D., Cahill, T.A., Dutcher, D., McDougall, E.M., & Stroud, K. (2001). Comparison of real-time instruments used
to monitor airborne particulate matter. Journal of Air and Waste Management Association, 51, 109–120.
Dedrick, D.E., Beyer, E.W., Rader, D.J., Klebanoff, L.E., & Leung, A.H. (2005). Verification studies of thermophoretic protection for extreme ultraviolet masks.
Journal of Vacuum Science & Technology B, 23 1071–1023.
Derjaguin, B.V., Rabinovich, Ya. I., Storozhilova, A.I., & Scherbina, G.I. (1976). Measurement of the coefficient of thermal slip of gases and the
thermophoresis velocity of large-size aerosol particles. Journal of Colloid Interface Science, 57, 451–461.
Dockery, D.W., Pope, C.A., Xu, X., Spengler, J.D., Ware, J.H., Fay, M.E., Ferris, B.G., Jr, & Speizer, F.E. (1993). An association between air pollution and
mortality in six U.S. cities. New England Journal of Medicine, 329, 1753–1759.
Geiser, M., Rothen-Rutishauser, B., Kapp, N., Schürch, S., Kreyling, W., Schulz, H., Semmler, M., Im Hof, V., Heyder, J., & Gehr, P. (2005). Ultrafine particles
cross cellular membranes by nonphagocytic mechanisms in lungs and in cultured cells. Environmental Health Perspectives, 113, 1555–1560.
Gonzalez, D., Nasibulin, A.G., Baklanov, A.M., Shandakov, S.D., Brown, D.P., Queipo, P., & Kauppinen, E.I. (2005). A new thermophoretic precipitator for
collection of nanometer-sized aerosol particles. Aerosol Science and Technology, 39, 1064–1071.
Green, H. L. and Watson, H. H.. (1935). Medical Research Council Special Report. no. 199, His Majesty’s Stationary Office, London.
Hinds, W.C. (1999). Aerosol Technology: Properties, Behavior, and Measurement of Airborne Particles. John Wiley & Sons, Inc: New York.
Holguin, F. (2008). Traffic, outdoor air pollution, and asthma. Immunology and Allergy Clinics of North America, 28, 577–588.
Lin, J.S., & Tsai, C.J. (2003). Thermophoretic deposition efficiency in a cylindrical tube taking into account developing flow at the entrance region. Journal of
Aerosol Science, 34, 569–583.
Lin, J.S., Tsai, C.J., & Chang, C.P. (2004). Suppression of particle deposition in tube flow by thermophoresis. Journal of Aerosol Science, 35, 1235–1250.
Katsouyanni, K., Touloumi, G., Spix, C., Schwartz, J., Balducci, F., Medina, S., Rossi, G., Wojtyniak, B., Sunyer, J., Bacharova, L., Schouten, J.P., Ponka, A., &
Anderson, H.R. (1997). Short term effects of ambient sulphur dioxide and particulate matter on mortality in 12 European cities: results from time
series data from the APHEA project. British Medical Journal, 314, 1658–1663.
Kethley, T.W., Gordon, M.T., & Orr, C., Jr. (1952). A thermal precipitator for aerobacteriology. Science, 116, 368–369.
Kim, S., Jaques, P.A., Chang, M.C., Barone, T., Xiong, C., Friedlander, S.K., & Sioutas, C. (2001). Versatile aerosol concentration enrichment system (VACES)
for simultaneous in vivo and in vitro evaluation of toxic effects of ultrafine, fine and coarse ambient particles Part I: Development and laboratory
characterization. Journal of Aerosol Science, 32, 1281–1297.
Lorenzo, R., Kaegi, R., Gehrig, R., Scherrer, L., Grobéty, B., & Burtscher, H. (2007). A thermophoretic precipitator for the representative collection of
atmospheric ultrafine particles for microscopic analysis. Aerosol Science and Technology, 41, 934–943.
Messerer, A., Niessner, R., & Pöschl., U. (2003). Thermophoretic deposition of soot aerosol particles under experimental conditions relevant for modern
diesel engine exhaust gas systems. Journal of Aerosol Science, 34, 1009–1021.
Montassier, N., Boulaud, D., & Renoux, A. (1991). Experimental study of thermophoretic particle deposition in laminar tube flow. Journal of Aerosol Science,
22, 677–687.
Orr, C., & Martin, R.A. (1958). Thermal precipitator for continuous aerosol sampling. Review of Scientific Instruments, 29, 129–130.
Patel, M.M., & Miller, R.L. (2009). Air pollution and childhood asthma: recent advances and future directions. Current Opinion in Pediatrics, 21, 235–242.
Pope, C.A., 3rd, & Dockery, D.W. (2006). Health effects of fine particulate air pollution: lines that connect. Air & Waste Management Association, 56,
709–742.
Reist, P.C. (1993). Aerosol Science and Technology (2nd ed.). McGraw-Hill: New York.
Salam, M.T., Islam, T., & Gilliland, F.D. (2008). Recent evidence for adverse effects of residential proximity to traffic sources on asthma. Current Opinion in
Pulmonary Medicine, 14, 3–8.
Santachiara, G, Prodi, F., & Cornetti, C. (2002). Experimental measurements on thermophoresis in the transition region. Journal of Aerosol Science, 33,
769–780.
Smith, J.N., Moore, K.F., Eisele, F.L., Voisin, D., Ghimire, A.K., Sakurai, H., & McMurry, P.H. (2005). Chemical composition of atmospheric nanoparticles
during nucleation events in Atlanta. Journal of Geophysical Research, 110, 1–13.
Tsai, C.J., & Lu, H.C. (1995). Design and evaluation of a plate-to-plate thermophoretic precipitator. Aerosol Science and Technology, 22, 172–180.
Tsai, C.J., Lin, J.S., & Aggarwal, S.J. (2004). Thermophoretic deposition of particles in laminar and turbulent tube flows. Aerosol Science and Technology, 38, 131–139.
Talbot, L., Cheng, R.K., Schefer, R.W., & Willis, D.R. (1980). Thermophoresis of particles in a heated boundary layer. Journal of Fluid Mechanics, 101, 737–758.
Wright, B.W. (1953). Gravimetric thermal precipitator. Science, 118, 195.

Das könnte Ihnen auch gefallen