25.01.2013 Views

A Design Tool for Aerodynamic Lens Systems - Department of ...

A Design Tool for Aerodynamic Lens Systems - Department of ...

A Design Tool for Aerodynamic Lens Systems - Department of ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Aerosol Science and Technology, 40:320–334, 2006<br />

Copyright c○ American Association <strong>for</strong> Aerosol Research<br />

ISSN: 0278-6826 print / 1521-7388 online<br />

DOI: 10.1080/02786820600615063<br />

A <strong>Design</strong> <strong>Tool</strong> <strong>for</strong> <strong>Aerodynamic</strong> <strong>Lens</strong> <strong>Systems</strong><br />

Xiaoliang Wang 1,2 and Peter H. McMurry 1<br />

1 <strong>Department</strong> <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Minnesota, Minneapolis, Minnesota, USA<br />

2 Current address: TSI Inc., St. Paul, Minnesota, USA<br />

We report in this article the development <strong>of</strong> a tool to design and<br />

evaluate aerodynamic lens systems: the <strong>Aerodynamic</strong> <strong>Lens</strong> Calculator.<br />

This Calculator enables quick and convenient design <strong>of</strong><br />

aerodynamic lens systems based on the parameterized knowledge<br />

gained from our detailed numerical simulations <strong>of</strong> flow and particle<br />

transport through aerodynamic lens systems. It designs the key<br />

dimensions <strong>of</strong> a lens system: pressure limiting orifice, relaxation<br />

chamber, focusing lenses, spacers and the accelerating nozzle. It<br />

also provides estimates <strong>of</strong> particle terminal axial velocities, particle<br />

beam width, and particle transmission efficiencies. This article<br />

describes in detail the in<strong>for</strong>mation that is used in the design tool,<br />

including equations <strong>for</strong> pressure drop through orifices, contraction<br />

factors as a function <strong>of</strong> Reynolds, Mach and Stokes numbers,<br />

spacer length as a function <strong>of</strong> Reynolds number and estimations<br />

<strong>for</strong> the relaxation chamber dimensions. The article also evaluates<br />

the per<strong>for</strong>mance <strong>of</strong> the design tool by comparing its predictions<br />

with the per<strong>for</strong>mance <strong>of</strong> aerodynamic lens systems that have been<br />

described in the literature.<br />

[Supplementary materials are available <strong>for</strong> this article. Go to the<br />

publisher’s online edition <strong>of</strong> Aerosol Science and Technology <strong>for</strong> the<br />

following free supplemental resources: An instruction manual <strong>for</strong><br />

the <strong>Aerodynamic</strong> <strong>Lens</strong> Calculator.]<br />

INTRODUCTION<br />

<strong>Aerodynamic</strong> lens systems have been widely used to generate<br />

collimated narrow particle beams since they were invented<br />

by Liu et al. (1995a, b). Typical applications include inlets <strong>for</strong><br />

aerosol mass spectrometers (Ziemann et al. 1995; Schreiner et al.<br />

1998, 1999, 2002; Jayne et al. 2000; Tobias et al. 2000; Öktem<br />

et al. 2004; Su et al. 2004; Svane et al. 2004; Drewnick et al.<br />

2005; Zelenyuk and Imre 2005), nanostructured material synthesis<br />

(Girshick et al. 2000; Dong et al. 2004; Piseri et al. 2004), and<br />

micro-scale device fabrication (Di Fonzo et al. 2000; Gidwani<br />

2003).<br />

In the original design <strong>of</strong> the aerodynamic lens system (Liu<br />

et al. 1995a, b), aerosols are sampled from one atmosphere pres-<br />

Received 23 November 2005; accepted 3 February 2006.<br />

This work was supported by NSF (Grant No. DMI-0103169)<br />

and the University <strong>of</strong> Minnesota Supercomputing Institute. We thank<br />

Dr. Frank Einar Kruis <strong>for</strong> helpful discussions.<br />

Address correspondence to Peter H. McMurry, 111 Church Street<br />

SE, Minneapolis, MN 55455, USA. E-mail: mcmurry@me.umn.edu<br />

320<br />

sure through a pressure limiting orifice into the lenses which<br />

operate at a pressure <strong>of</strong> ∼200 Pa, with low pressure drop across<br />

each lens. Typically a series <strong>of</strong> three to five lenses progressively<br />

focus spherical particles in the size range <strong>of</strong> 25–250 nm (near<br />

unit density) into a narrow beam. The collimated particle beam<br />

is then delivered through an accelerating nozzle to the detection<br />

chamber maintained at very low pressure. Most lens systems<br />

being used today are similar in design to the system described<br />

by Liu et al. (1995a, b).<br />

Several studies have reported on lens systems that were designed<br />

to operate under conditions that are significantly different<br />

from those used by Liu et al. (1995 a, b). For example, Di Fonzo<br />

et al. (2000) designed a lens system to focus silicon carbide<br />

particles in the diameter range <strong>of</strong> 10–100 nm using an argonhydrogen<br />

mixture; Dong et al. (2004) used a lens system to focus<br />

10–200 nm silicon particles in hydrogen. Schreiner et al. developed<br />

aerodynamic lens systems operating at elevated pressures<br />

(2.5–20 kPa) <strong>for</strong> stratospheric aerosol studies (Schreiner et al.<br />

1998, 1999, 2002). Lee et al. (2003) used a single lens to focus<br />

micron-sized particles in the atmospheric pressure range. Wang<br />

et al. (2005a, b) developed a systematic procedure to design<br />

aerodynamic lenses <strong>for</strong> 3–30 nm nanoparticles.<br />

To characterize the per<strong>for</strong>mance <strong>of</strong> an aerodynamic lens system,<br />

detailed numerical or experimental studies are required. Liu<br />

et al. (1995a, b) carried out theoretical, numerical, and experimental<br />

analysis <strong>of</strong> the per<strong>for</strong>mance <strong>of</strong> aerodynamic lens systems.<br />

Zhang et al. (2002, 2004) repeated the simulations by Liu<br />

et al. with a compressible flow model. To study particle loss and<br />

beam broadening due to diffusion, Gidwani (2003) and Wang<br />

et al. (2005b) simulated the flow and particle transport in the<br />

lens system considering the particle Brownian motion.<br />

Although numerical simulations or experimental evaluation<br />

<strong>of</strong> the per<strong>for</strong>mance <strong>of</strong> lens systems can provide more accurate<br />

results, they are very costly in both time and financial resources.<br />

Furthermore, these studies require predefined dimensions and<br />

operating conditions <strong>of</strong> the aerodynamic lens system, which<br />

need to be iteratively optimized. Due to the wide applications <strong>of</strong><br />

aerodynamic lenses, a simple tool that can provide quick and reasonably<br />

accurate lens design and evaluation would be useful. We<br />

report the development <strong>of</strong> such an <strong>Aerodynamic</strong> <strong>Lens</strong> Calculator<br />

in this article. We studied flow and particle transport through<br />

a single lens at a wide range <strong>of</strong> Reynolds and Mach numbers.


From these simulations, we obtained in<strong>for</strong>mation about particle<br />

contraction factors, transport efficiencies, flow discharge coefficients<br />

through orifices and the lengths <strong>of</strong> spacer <strong>for</strong> flow to<br />

redevelop. We also simulated flow through the pressure limiting<br />

orifice and the relaxation region, the accelerating nozzle, as<br />

well as the entire lens system (Wang 2005b). Results from these<br />

detailed numerical calculations were parameterized and incorporated<br />

into the design tool described in this paper. We correlated<br />

the particle terminal axial velocity in the vacuum chamber<br />

downstream <strong>of</strong> the nozzle as a function <strong>of</strong> Stokes number with a<br />

simple equation. This enabled us to estimate the particle terminal<br />

axial velocity under similar nozzle geometry and operating<br />

pressure. We also provide estimate <strong>of</strong> the particle beam width<br />

both inside and downstream <strong>of</strong> the lens system. Both aerodynamic<br />

focusing and diffusion broadening are taken into account<br />

in the beam width estimation. We considered three particle loss<br />

mechanisms to calculate the particle transport efficiency: inertia<br />

impaction on the orifice plate, impaction on the spacer walls due<br />

to defocusing, and losses due to diffusion.<br />

This design tool avoids the need <strong>for</strong> the time consuming detailed<br />

numerical simulations, and enables designers to quickly<br />

test a range <strong>of</strong> design options. The tool calculates the lens dimensions,<br />

operating conditions and flow parameters at each<br />

lens stage. It also estimates the terminal particle velocity after<br />

the choked accelerating nozzle, particle beam diameter at each<br />

stage and at the location <strong>of</strong> the detector, and particle transport<br />

efficiency through the lens system. This article describes the detailed<br />

in<strong>for</strong>mation used in the Calculator. We first describe the<br />

analytical, empirical or numerical relations used to design the<br />

dimensions and operating conditions <strong>of</strong> aerodynamic lens systems.<br />

Next we describe the method used to estimate the particle<br />

terminal velocity, beam width and transport efficiency. Finally,<br />

we report the validation <strong>of</strong> this lens design tool by comparing<br />

its predictions with several detailed numerical and experimental<br />

studies reported in the literature.<br />

AERODYNAMIC LENS SYSTEM DESIGN<br />

The objective <strong>of</strong> aerodynamic lens design is to build a<br />

lens system that has best per<strong>for</strong>mance, i.e., maximum particle<br />

focusing, minimum particle losses, and minimum pumping<br />

A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 321<br />

FIG. 1. Schematic and nomenclatures <strong>of</strong> the lens system.<br />

requirements under user specified inputs (particle size range,<br />

particle density, etc.). To achieve this goal, one needs to optimize<br />

dimensions <strong>of</strong> the lens system (number <strong>of</strong> lenses, lens diameters,<br />

inner diameter and length <strong>of</strong> the spacers between two<br />

lenses, the flow limiting orifice, relaxation chamber and the accelerating<br />

nozzle) and operating parameters (flowrate, pressure,<br />

and carrier gas).<br />

We have briefly described the aerodynamic lens design procedure<br />

in an earlier article with emphasis on lenses <strong>for</strong> nanoparticles<br />

(Wang et al. 2005a). In this section, we describe the design<br />

principle in more detail, and generalize the guidelines to be applicable<br />

to particles ranging in size from several nanometers to<br />

greater than one micrometer.<br />

Assumptions<br />

Figure 1 shows a schematic <strong>of</strong> an aerodynamic lens system<br />

with the nomenclature used in this paper and in the code <strong>of</strong> the<br />

lens calculator. The lens system described in this article consists<br />

<strong>of</strong> the following parts: an inlet orifice followed by a relaxation<br />

chamber, one or more lenses separated by spacers, and an accelerating<br />

nozzle. The pressure limiting orifice determines the<br />

volumetric flowrate through the lens system, and the accelerating<br />

nozzle defines the operating pressure. We assume that the<br />

lenses are primarily responsible <strong>for</strong> all focusing. The focusing<br />

or defocusing effects that might be caused by the inlet orifice<br />

or the accelerating nozzle are not controlled by the design tool,<br />

but the user can optimize these effects by varying the operating<br />

pressure. The inner diameters <strong>of</strong> each spacer are usually the<br />

same so as to simplify machining.<br />

Particles are assumed to be spherical and electrically neutral.<br />

The effects <strong>of</strong> particle shape have been discussed by Liu et al.<br />

(1995a, b) and Huffman et al. (2005). The design tool restricts<br />

flow through the lenses to be laminar, subsonic, and continuum.<br />

It is well known that flow instability and turbulence can disperse<br />

particles and destroy focusing. However, the Reynolds<br />

number above which these effects contribute to defocusing is not<br />

known with certainty. Eichler et al. (1998) and Gómez-Moreno<br />

et al. (2002) reported that turbulent transition in an orifice flow<br />

occurred at Reynolds number around 70. Back and Roschke<br />

(1972) and Gong et al. (1996) found flow instabilities around


322 X. WANG AND P. H. MCMURRY<br />

aReynolds number <strong>of</strong> 200 downstream <strong>of</strong> sudden expansions.<br />

Our design tool constrains Reynolds numbers (Re) tobebelow<br />

200. There<strong>for</strong>e,<br />

Re = ρ1ud f<br />

µ = 4 ˙m<br />

πµd f<br />

≤ 200. [1]<br />

Although sonic or supersonic nozzles have been widely used<br />

to focus particles (Murphy and Sears 1964; Israel and Friedlander<br />

1967; Cheng and Dahneke 1979; Dahneke and Cheng 1979;<br />

Fernández de la Mora et al. 1989; Mallina et al. 2000; Tafreshi<br />

et al. 2002), we constrain the flow through lenses to be subsonic<br />

to avoid the complicated influence <strong>of</strong> shock waves on particle<br />

focusing. There<strong>for</strong>e, the lens Mach number (Ma) islimited to<br />

be smaller than the value corresponding to choked flow (Ma ∗ )<br />

(Wang et al. 2005a), i.e.,<br />

Ma = u<br />

c < Ma∗ . [2]<br />

If the flow is in the free molecular regime, the particle inertia<br />

would greatly exceed the drag <strong>for</strong>ce and no focusing could be<br />

achieved. Furthermore, rarefied gas dynamics is too complicated<br />

to handle using our simple design tool. There<strong>for</strong>e, we restrict the<br />

flow Knudsen number (Kn)tobesmaller than 0.1 to assume that<br />

the flow is continuum,<br />

Kn = 2λ1<br />

d f<br />

< 0.1. [3]<br />

Focusing <strong>Lens</strong><br />

The parameter that governs particle focusing through an aerodynamic<br />

lens is the Stokes number (St), which is defined as the<br />

ratio <strong>of</strong> the particle stopping distance at the average orifice velocity<br />

(u) tothe orifice diameter (d f ):<br />

St = τu<br />

d f<br />

= 2ρpd 2 pCc ˙m<br />

9πρ1µd3 , [4]<br />

f<br />

where Cc = 1 + Knp(1.257 + 0.4e −1.1<br />

Knp ) and Knp = 2λ1/dp.<br />

As shown in earlier studies, the particle contraction factor (ηc),<br />

which is the ratio <strong>of</strong> terminal to initial radial positions <strong>of</strong> the<br />

particles travelling through the lens, is a strong function <strong>of</strong> the<br />

Stokes number: |ηc| ≈1 when St ≪ 1, |ηc| < 1 when St ≈ 1,<br />

and |ηc| > 1 when St ≫ 1. In other words, small particles<br />

follow gas streamlines and thus are not focused, intermediatesized<br />

particles are focused and large particles are defocused by<br />

the lens. There exists an optimum Stokes number (Sto) <strong>for</strong> which<br />

ηc = 0 (Liu et al. 1995a). Liu et al. showed that the contraction<br />

factors are independent <strong>of</strong> the particle initial radial position <strong>for</strong><br />

particles that are not too far away from the axis. We refer to these<br />

near-axis contraction factors unless stated otherwise.<br />

By rearranging Equation (4), we can calculate the diameter<br />

<strong>of</strong> the lens aperture <strong>for</strong> optimally focused particles:<br />

�<br />

2ρpd<br />

d f =<br />

2 pCc � 1<br />

˙m 3<br />

. [5]<br />

9πρ1µSto<br />

Note that the optimal Stokes number is a function <strong>of</strong> flow<br />

Reynolds number, Mach number, and lens geometry (Liu et al.<br />

1995a; Zhang et al. 2002). We use thin plate orifices as the focusing<br />

elements in this paper. Liu et al. (1995a) showed that<br />

the dependence <strong>of</strong> Sto on the constriction ratio (β)<strong>of</strong>thin plate<br />

orifices becomes weak when β ≤ 0.25. To simplify the design<br />

process, we use β ≤ 0.25 in the design tool. In principle, one<br />

could use β>0.25 if the ηc vs. St curves were known.<br />

To obtain the relation between Sto, Ma, and Re, wehave<br />

carried out numerical simulations <strong>of</strong> particle motion through a<br />

single lens at a range <strong>of</strong> Mach number (Ma = 0.03–0.32) and<br />

Reynolds numbers (Re = 1–100) which cover the typical operating<br />

conditions <strong>of</strong> lenses. In these simulations, β was set equal<br />

to 0.2. Particles were injected at an initial radial location (r pi)<strong>of</strong><br />

0.15ds from the lens axis and Brownian motion was neglected.<br />

The relationship between the contraction factor and St, Re, and<br />

Ma is shown in Figure 2(a)–(c). Note that the dependence <strong>of</strong><br />

contraction on Reynolds number decreases as Mach number<br />

increases. The dependence <strong>of</strong> the contraction factor on the particle<br />

initial radial location becomes significant <strong>for</strong> large Stokes<br />

numbers (Liu et al. 1995a). There<strong>for</strong>e, the near-axis contraction<br />

factors in Figure 2 are not very representative at the higher St end<br />

<strong>of</strong> each curve, and significant impaction losses occur <strong>for</strong> particles<br />

with St > 10. Figure 3 shows the optimum Stokes number<br />

Sto as a function <strong>of</strong> Re <strong>for</strong> three different Ma, which corresponds<br />

to the St where ηc = 0ateach curve in Figure 2. The value <strong>of</strong><br />

Sto <strong>for</strong> a specific set <strong>of</strong> Re and Ma can be interpolated from this<br />

data.<br />

From Equation (5), we can see that the pressure upstream<br />

<strong>of</strong> the lens is an important parameter in calculating the lens<br />

diameter. There<strong>for</strong>e, we need to accurately estimate the pressure<br />

drop across an orifice. We have developed a model <strong>for</strong> viscous<br />

flow through an orifice<br />

CdY �<br />

˙m = A f � 2ρ1(p1 − p2)<br />

1 − β4 CdY<br />

= A f �<br />

1 − β4 p1<br />

�<br />

2M �p<br />

RT1 p1<br />

as well as <strong>for</strong> the discharge coefficient (Wang et al. 2005a)<br />

Cd =<br />

⎧<br />

0.1373Re<br />

⎪⎨<br />

⎪⎩<br />

0.5 (Re < 12)<br />

1.118 − 0.8873 × Ln(Re) + 0.3953 × [Ln(Re)] 2<br />

− 0.07081 × [Ln(Re)] 3<br />

+ 0.005551 × [Ln(Re)] 4 − 0.0001581 × [Ln(Re)] 5<br />

(12 ≤ Re < 5000)<br />

0.59 (Re ≥ 5000) [7]<br />

[6]


FIG. 2. Near-axis particle contraction factor as a function <strong>of</strong> the Stokes number<br />

<strong>for</strong> various Reynolds numbers at three different subsonic Mach numbers.<br />

(a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.<br />

A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 323<br />

FIG. 3. Optimum Stokes number as a function <strong>of</strong> Reynolds number <strong>for</strong> three<br />

different Mach numbers.<br />

The expansion factor Y is calculated using the expression <strong>of</strong><br />

(Bean 1971):<br />

Y = 1 − (0.410 + 0.350β 4 ) �p<br />

. [8]<br />

γ p1<br />

Using Equations (5–8) we can calculate the diameter <strong>of</strong> each<br />

lens <strong>for</strong> a given mass flowrate and a pressure at any stage <strong>of</strong> the<br />

lens assembly.<br />

To study the particle size range that a multiple lens system<br />

can focus, we found from Equation (4) that in the free molecular<br />

regime St is proportional to dp while in the continuum regime<br />

it is proportional to d2 p .Ifalens system focuses particles <strong>of</strong> a<br />

decade size range dp1–dpN with dp1 = 10dpN, then the Stokes<br />

number <strong>of</strong> dp1 at the last lens will be 10Sto if it is in the free<br />

molecular regime, and 100Sto if it is in the continuum regime. If<br />

by any non-ideal effects particles <strong>of</strong> dp1were not focused exactly<br />

onto the axis, they will start to defocus at the last lens or even at<br />

an earlier stage. There<strong>for</strong>e, it is probably a good idea to design<br />

a lens system to cover approximately a decade <strong>of</strong> particle size.<br />

Substantial focusing can be achieved even with sub-optimal<br />

Stokes numbers when a sufficient number <strong>of</strong> lenses are used.<br />

There are two strategies to focus a wider range <strong>of</strong> particle sizes.<br />

First, one can add additional lenses to refocus the larger particles<br />

be<strong>for</strong>e they are defocused and lost. Second, one can use<br />

first several lenses to focus the largest particles (dp1) toagiven<br />

tolerance close to the axis, and then design the following lenses<br />

at ηc(dp1) =−1 and focus smaller sizes sub-optimally.<br />

If particles are large enough that diffusion is not significant,<br />

and the largest particles do not defocus in the final stages,<br />

the more lenses used in an assembly, the better focusing can<br />

be achieved. In practice, however, increasing the number <strong>of</strong><br />

lenses increases the difficulty <strong>of</strong> alignment and adds size and expense.<br />

Effects <strong>of</strong> diffusion increase as the lens length increases.


324 X. WANG AND P. H. MCMURRY<br />

Furthermore, larger particles are more prone to be defocused in<br />

an assembly with more lenses. Typically, we have found that a<br />

lens assembly consisting <strong>of</strong> five lenses is sufficient to focus particles<br />

<strong>of</strong> a decade size range. Three or four lenses were used in<br />

our previous study <strong>of</strong> nanoparticle focusing (Wang et al. 2005b).<br />

Operating Pressures<br />

From Equation (4), we obtain the pressure (pfocusing) <strong>for</strong> focusing<br />

particles dp using the ideal gas law<br />

pfocusing = 2ρpd 2 p Cc ˙mRT<br />

9πµd 3 f MSto<br />

. [9]<br />

Note that Cc is a function <strong>of</strong> pfocusing. The minimum pressure <strong>for</strong><br />

flow to be subsonic (pMa) and continuum (pKn) are respectively<br />

given as (Wang et al. 2005a)<br />

and<br />

˙mc<br />

pMa =<br />

CdYc A f<br />

pKn =<br />

2 T1<br />

λr<br />

d f Kn∗ Tr<br />

� �<br />

2M<br />

pr<br />

RT1<br />

xc<br />

[10]<br />

1 + S/Tr<br />

. [11]<br />

1 + S/T1<br />

Figure 4 shows examples <strong>of</strong> pMa, pKn, and the required pfocusing<br />

to focus particles <strong>of</strong> different sizes as functions <strong>of</strong> the orifice size<br />

in Figure (4a) air and Figure (4b) helium when the flowrate is 0.1<br />

standard liter per minute (slm). The lens operating pressures are<br />

those along the pfocusing curves where pfocusing is wider than both<br />

pMa and pKn. Note that larger particles have a larger operating<br />

pressure window, and that using a lighter carrier gas can increase<br />

the operating pressures <strong>for</strong> smaller sizes which enables focusing<br />

them. Also shown in Figure 4(a) is a “Re warning line,” which<br />

corresponds to Re = 200 above which turbulence is likely to<br />

degrade focusing.<br />

As was pointed out previously (Wang et al. 2005a), increasing<br />

the operating pressure can reduce the detrimental effects <strong>of</strong><br />

nanoparticle diffusion and can reduce the pumping needs. From<br />

Figure 4, we can see that the maximum operating pressure <strong>of</strong><br />

smaller sizes occurs where the pMa curve intersects the pfocusing<br />

curve <strong>for</strong> each particle size, and its value can be derived from<br />

Equations (9) and (10),<br />

pmax = 648 ˙mµ2 St 2 o<br />

πC 3 d Y 3 c ρ2 p d4 p C2 c<br />

�<br />

M<br />

2RTx 3 c<br />

. [12]<br />

Note that this is an implicit expression because Cc is a function<br />

<strong>of</strong> pmax, and an iterative method is needed to calculate pmax.<br />

The pMa curve may not intersect the pfocusing curve <strong>for</strong> larger<br />

particles in the practical orifice diameter range. In this case any<br />

value <strong>of</strong> pfocusing larger than pKn can be used. However, Figure 4<br />

FIG. 4. The operating pressures <strong>for</strong> focusing unit density particles <strong>of</strong> different<br />

sizes as a function <strong>of</strong> orifice size using (a) air and (b) helium as the carrier gas.<br />

The flowrate is 0.1 slm. The two dash lines are the lower pressure limits pMa<br />

and pKn, respectively. The solid lines are pfocusing <strong>for</strong> indicated sizes.<br />

shows that the higher the operating pressure, the larger is the<br />

Reynolds number. It follows that Reynolds numbers tend to limit<br />

the maximum pressure at which lenses can be operated.<br />

Length <strong>of</strong> Spacers<br />

Spacers are used to separate lenses. They provide room <strong>for</strong><br />

generating the periodic converging/diverging flow pattern with<br />

orifices, which drives aerodynamic focusing. In addition to the<br />

constraints mentioned earlier (β ≤ 0.25; equal inner diameter <strong>of</strong><br />

all spacers), the spacers should be long enough so that the flow<br />

from the preceding lens can relax back to fully developed pipe<br />

flow be<strong>for</strong>e reaching the next lens (Liu et al. 1995b). Typically,<br />

spacer lengths are 10–50 times the upstream orifice diameter<br />

depending on the orifice Reynolds number (Wang et al. 2005a).<br />

However, no quantitative guideline on the spacer length has been<br />

reported.


Figure 5 shows the flow field through two lenses separated<br />

by a spacer. The flow separates from the wall when it passes<br />

through the lens and a recirculation zone <strong>for</strong>ms downstream<br />

<strong>of</strong> the orifice. After a distance lr (reattachment length), the flow<br />

reattaches to the wall, and becomes fully developed at a distance<br />

<strong>of</strong> l f (redevelopment length) downstream <strong>of</strong> the orifice. The<br />

existence <strong>of</strong> the downstream lens is felt by the flow la (approach<br />

length) upstream <strong>of</strong> the next lens where the flow starts to curve<br />

toward the centerline. Ideally, the minimum length <strong>of</strong> spacers<br />

(ls) should be ls1 = lr + la, and more safely, ls2 = l f + la.<br />

If the spacer length is less than ls1, the recirculation zone will<br />

likely fill the whole length <strong>of</strong> the spacer and less focusing will<br />

occur. Furthermore, particles are more likely to be entrained in<br />

the recirculation zones and be lost to the wall. There<strong>for</strong>e, when<br />

particle diffusion is negligible and the overall length <strong>of</strong> the lens<br />

system is not a concern, we suggest using spacer lengths at<br />

least ls2. When diffusion is important, ls1 < ls ≤ ls2 can be<br />

used.<br />

To estimate the values <strong>of</strong> lr, l f , and la,wehave carried out numerical<br />

simulations <strong>of</strong> flow through single lenses in the Reynolds<br />

number range <strong>of</strong> 0.1–200. Calculated values <strong>for</strong> lr are shown in<br />

Figure 6, together with data provided by other researchers (Back<br />

and Roschke 1972; Oliveira et al. 1998). Note that our numerical<br />

results are very close to those by Oliveria (1998), which apply<br />

to similar Reynolds numbers and expansion ratios. The results<br />

by Back et al. differ somewhat from ours and those <strong>of</strong> Oliveria,<br />

probably due to the difference in ER. Since we did not carry out<br />

simulations <strong>for</strong> Re > 200, we will use the data by Back et al.<br />

to estimate lr in this Re range, noting that applying Back’s data<br />

(ER = 2.6) to larger ER values might lead to some error. Also<br />

shown in Figure 6 is a fitted curve to the whole Reynolds number<br />

range, which can be described as follows<br />

lr<br />

h =<br />

⎧<br />

−1.016 × 10<br />

⎪⎨<br />

⎪⎩<br />

−6 Re3 + 3.150 × 10−4 Re2 + 8.407<br />

× 10−2 Re + 0.3689<br />

�<br />

<strong>for</strong> Re ≤ 200<br />

9.096 + 3860.168 × exp − log10 (Re)<br />

�<br />

.[13]<br />

�<br />

� 0.431<br />

× sin − 4.226 <strong>for</strong> Re > 200.<br />

2π log 10 (Re)<br />

1.272<br />

A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 325<br />

FIG. 5. Streamlines through two lenses separated by a spacer.<br />

The redevelopment length l f can be estimated by l f = lr + le<br />

where le is given as (Young et al. 2000)<br />

le<br />

dt<br />

�<br />

0.06Res <strong>for</strong> 0 ≤ Res < 2300<br />

=<br />

. [14]<br />

<strong>for</strong> Res ≥ 2300<br />

4.4Re 1<br />

6<br />

s<br />

The approach length la is found in our simulation to be independent<br />

<strong>of</strong> Reynolds number (Re < 150) with<br />

la/h ≈ 1.625. [15]<br />

Since la is relatively shorter than lr and l f ,weuse the above<br />

equation to estimate la <strong>for</strong> the whole Reynolds number range.<br />

Flow Limiting Orifice, Relaxation Chamber,<br />

and Accelerating Nozzle<br />

The flow limiting orifice, relaxation chamber, and the accelerating<br />

nozzle are important components <strong>of</strong> the aerodynamic<br />

FIG. 6. Reattachment length <strong>of</strong> flow downstream <strong>of</strong> an axisymmetric<br />

expansion.


326 X. WANG AND P. H. MCMURRY<br />

lens system. The flow limiting orifice and relaxation chamber<br />

are required if the particle source pressure is higher than the<br />

lens operating pressure. The orifice sets the volumetric flowrate<br />

through the lens system, and reduces the pressure to the lens<br />

operating pressure. Usually the flow through the flow limiting<br />

orifice is critical. There<strong>for</strong>e, a relaxation chamber is needed to<br />

slow down the high velocity flow and particles to prevent particle<br />

impaction on the downstream lens. The relaxation chamber<br />

also provides space <strong>for</strong> the flow to reattach to the wall after the<br />

recirculation eddies downstream <strong>of</strong> the orifice. The accelerating<br />

nozzle controls the exact lens operating pressure, and it accelerates<br />

particles to a downstream destination with minimized<br />

divergence angles.<br />

The inner diameter <strong>of</strong> the relaxation chamber can be estimated<br />

from the radial stopping distance <strong>of</strong> the largest particles<br />

<strong>of</strong> interest. The lower half <strong>of</strong> Figure 7(a) shows the flow streamlines<br />

downstream <strong>of</strong> a straight-bored critical orifice with a diameter<br />

and wall thickness <strong>of</strong> 0.1 mm. The inlet pressure is 1<br />

atm and the outlet pressure is 266 Pa. The flowrate is 70.6 sccm,<br />

which correspond to a Reynolds number <strong>of</strong> 1071 based on the<br />

orifice diameter. There<strong>for</strong>e, the jet is turbulent. However, <strong>for</strong> the<br />

sake <strong>of</strong> simplicity, we only used a steady laminar flow model<br />

in this simulation and hence the streamlines in Figure 7(a) only<br />

serve as a qualitative description <strong>of</strong> the averaged flow behavior.<br />

The upper half <strong>of</strong> Figure 7(a) shows the half jet opening angle<br />

θ, which is defined as θ = tan −1 (h/lr) ≈ tan −1 (vr/va). Assuming<br />

va equals the speed <strong>of</strong> sound c, then vr ≈ c tan(θ). We<br />

further conservatively assume that particle velocity equals to the<br />

FIG. 7. Straight bored critical orifice and cylindrical relaxation chamber.<br />

(a) Flow streamlines and illustration <strong>of</strong> the half jet opening angle; (b) Trajectories<br />

<strong>of</strong> 100 nm (above axis) and 1 µm (below axis) particles.<br />

flow velocity, then the particle radial stopping distance can be<br />

estimated from Sr = τvr where τ is the relaxation time <strong>of</strong> the<br />

largest particles in the target focusing size range downstream <strong>of</strong><br />

orifice. The minimum diameter <strong>of</strong> the relaxation chamber (dr)<br />

is then<br />

dr = 2Sr + d f . [16]<br />

Here d f is the diameter <strong>of</strong> the critical orifice.<br />

The relaxation chamber should be long enough to let flow<br />

reattach and particles slow down. The length l f <strong>for</strong> flow to redevelop<br />

can be calculated using Equations (13) and (14). Due to<br />

the complicated shock structure, we assume that particles have<br />

velocities equal to the speed <strong>of</strong> sound c <strong>for</strong> the carrier gas at a<br />

distance lr downstream <strong>of</strong> the orifice while the gas velocity is<br />

negligibly low. The particle axial stopping distance downstream<br />

<strong>of</strong> lr is then Sa = τc. There<strong>for</strong>e, the length <strong>of</strong> the relaxation<br />

chamber Lr can be estimated as<br />

Lr = max(l f + la, lr + Sa). [17]<br />

The shape <strong>of</strong> the relaxation chamber and critical orifice aperture<br />

are also important to reduce particle losses. Figure (7b)<br />

shows trajectories <strong>of</strong> 100 nm (above the axis) and 1 µm (below<br />

the axis) particles through the critical orifice. The turbulent<br />

dispersion <strong>of</strong> particles is not included in the simulation. Note<br />

that some particles <strong>of</strong> both sizes are trapped inside the recirculation<br />

region <strong>for</strong> a long time, which makes them susceptible<br />

to coagulation or deposition losses. There<strong>for</strong>e, we recommend<br />

that the strong recirculation region be eliminated. This<br />

can be achieved by inserting a conical expansion downstream<br />

<strong>of</strong> the orifice as shown in Figure 8(a). Figure 8(b) shows trajectories<br />

<strong>of</strong> 100 nm and 1 µm particles through this modified<br />

relaxation chamber. Although this new design eliminates<br />

the recirculation eddy, impaction losses <strong>for</strong> larger particles increase<br />

due to the narrowed flow path, as is illustrated by the<br />

1 µm trajectories. To reduce impaction losses <strong>of</strong> larger particles,<br />

one can use conical nozzles, short capillaries or their<br />

combinations instead <strong>of</strong> thin plate orifices, since these nozzle<br />

shapes will reduce particle accelerations at the nozzle inlet<br />

(Fernández de la Mora and Riesco-Chueca 1988; Fernández de<br />

la Mora et al. 1989). Figure (8c) illustrates that impaction losses<br />

<strong>of</strong> 1 µm particles are reduced by adding a 0.1 mm thick 45 ◦<br />

chamfer to the original orifice. (Note that this orifice modification<br />

increases the flowrate by 10.4%.) Although we demonstrated<br />

methods to improve the per<strong>for</strong>mance <strong>of</strong> the relaxation<br />

region in this section, we should point out that more careful<br />

studies <strong>of</strong> the orifice shape and chamber dimensions are still<br />

required.<br />

We follow the recommendation <strong>of</strong> Liu et al. (1995b) to use<br />

a step nozzle as the accelerating nozzle <strong>of</strong> lens systems (see<br />

Figure 1). The diameter <strong>of</strong> the final orifice in the nozzle can<br />

be calculated using Equation (6), and other dimensions <strong>of</strong> the<br />

nozzle are taken as dt ≈ 2dn, Lt ≈ ds.Again, further research is


FIG. 8. Modified relaxation chamber with a conical divergent section to reduce<br />

recirculation and particle loss. (a) Flow streamlines (straight orifice); (b)<br />

Trajectories <strong>of</strong> 100 nm (above axis) and 1 µm (below axis) particles (straight<br />

orifice); (c) Trajectories <strong>of</strong> 100 nm (above axis) and 1 µm (below axis) particles<br />

(orifice with a chamfer).<br />

needed to optimize the nozzle design so that the nozzle can also<br />

achieve maximum focusing <strong>for</strong> a given size range and operating<br />

conditions.<br />

LENS PERFORMANCE ESTIMATION<br />

The most important per<strong>for</strong>mance parameters <strong>for</strong> an aerodynamic<br />

lens system are: particle terminal velocities, particle beam<br />

widths at various locations, and the particle transmission efficiencies.<br />

These parameters need to be evaluated by detailed numerical<br />

simulations, and ultimately by experiments. However,<br />

it is attractive to have a best estimation <strong>of</strong> the lens per<strong>for</strong>mance<br />

with a “one-button-click” ef<strong>for</strong>t during the design process. In<br />

this section, we describe the method used in the <strong>Lens</strong> Calculator<br />

to do the estimations.<br />

A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 327<br />

FIG. 9. Normalized particle terminal axial velocities downstream <strong>of</strong> several<br />

aerodynamic lens systems.<br />

Estimation <strong>of</strong> Particle Terminal Velocities<br />

We assume that the pressure downstream <strong>of</strong> the accelerating<br />

nozzle is low enough so that the gas-particle collisions are<br />

negligible and particles achieve their “terminal” velocities. We<br />

further assume that particles are in thermal equilibrium with the<br />

carrier gas upstream <strong>of</strong> the nozzle exit and their radial velocity<br />

can be approximately described by the Maxwell-Boltzmann distribution.<br />

This velocity distribution can be assumed to be frozen<br />

during the expansion downstream <strong>of</strong> the nozzle (Liu et al. 1995a;<br />

Wang et al. 2005b). There<strong>for</strong>e, the terminal radial velocity vpr<br />

can be estimated as<br />

f (vpr)dvpr = m �<br />

p<br />

exp −<br />

2πkTpF<br />

m pv2 �<br />

pr<br />

2π vpr dvpr. [18]<br />

2kTpF<br />

The particle terminal axial velocities (u p) depend on particle<br />

size, shape, nozzle geometry, pressure ratio, and carrier gases<br />

(Cheng and Dahneke 1979; Dahneke and Cheng 1979; Mallina<br />

et al. 1997). Figure 9 summarizes the particle terminal axial velocities<br />

<strong>for</strong> the four aerodynamic lens systems listed in Table 1.<br />

These four lens systems differ in dimensions, flowrate, pressure,<br />

and carrier gas. But they all use step accelerating nozzles similar<br />

to that described by Liu et al. (1995b), and the pressures downstream<br />

<strong>of</strong> the nozzles are less than 1 Pa. Note that the following<br />

equation fits the data points reasonably well in the St = τ c/dn<br />

range where data is available (0.01 < St < 80)<br />

u p/c = (0.939 + 0.09St)/(1 + 0.543St). [19]<br />

We will use this equation to estimate particle axial velocity in<br />

the <strong>Lens</strong> Calculator.


328 X. WANG AND P. H. MCMURRY<br />

TABLE 1<br />

Key features <strong>of</strong> four lens assemblies with step accelerating nozzles used in particle axial velocity comparison<br />

<strong>Lens</strong> system dn (mm) ds (mm) dt (mm) Lt (mm) Gas p1 (Pa) dp (nm)<br />

A(Wang et al. 2005b) 2.76 10 6 5 He 528 3–30<br />

B (Liu et al. 1995b) 3.0 10 6 10 Air 93–333 25–250<br />

C (TSI AFL-050) 3.2 17.78 7.62 17.78 Air 250 50–500<br />

D (TSI AFL-100) 1.48 17.78 7.62 17.78 Air 600 100–3000<br />

Estimation <strong>of</strong> Particle Beam Widths<br />

The particle beam width is defined as the beam diameter that<br />

encloses 90% <strong>of</strong> the total particle flux. In this section expressions<br />

are given that are used in the Calculator to calculate particle beam<br />

widths downstream <strong>of</strong> each lens as well as downstream <strong>of</strong> the<br />

accelerating nozzle. Bean widths are controlled by two factors:<br />

aerodynamic focusing and diffusion broadening. The focusing<br />

can be inferred from Figure 2, and the diffusion broadening<br />

can be estimated by the root mean square displacement xrms √ =<br />

2Dt, where D is the particle diffusion coefficient, and t is the<br />

particle residence time.<br />

Assuming the diameters <strong>of</strong> spacers upstream and downstream<br />

<strong>of</strong> the pressure limiting orifice (lens 0) are the same, the flow<br />

is fully developed upstream <strong>of</strong> the pressure limiting orifice, and<br />

particles are homogenously distributed in the flow cross section,<br />

the starting particle beam diameter enclosing 90% flux is then<br />

∼0.827ds,0. The particle beam diameter be<strong>for</strong>e reaching lens 1<br />

is<br />

dB,0 = 0.827ds,0ηc,0 + xrms,0. [20]<br />

The beam diameter dB,i at stage i inside the lens system can be<br />

easily calculated as<br />

dB,i = dB,i−1ηc,i + xrms, i, i = 1ton. [21]<br />

However, we should note that dB,i may exceed the spacer diameter<br />

ds,i due to defocusing or diffusion. In that case, we assume<br />

all particles outside the beam diameter <strong>of</strong> ds,i are lost and reset<br />

dB,i = 0.827 ds,i.<br />

Assuming particles achieve a Maxwell-Boltzmann distribution<br />

<strong>of</strong> radial velocities downstream <strong>of</strong> the acceleration nozzle<br />

(lens n + 1), we can estimate the particle beam width at a distance<br />

L downstream <strong>of</strong> the nozzle as follows:<br />

dB,n+1 = dB,nηc,n+1 + 3.04<br />

�<br />

2kTpF<br />

m p<br />

L<br />

. [22]<br />

Note that ηc,n+1is only a very rough estimate because the pressure<br />

downstream <strong>of</strong> the nozzle is so low that the definition <strong>of</strong><br />

contraction factor is no longer valid.<br />

u p<br />

Estimation <strong>of</strong> Particle Transmission Efficiencies<br />

Particle losses in an aerodynamic lens system arise from impaction<br />

and diffusion. Particle impaction happens both on the<br />

orifice plate and on the spacer walls.<br />

We can view the orifices (including the pressure limiting<br />

orifice, lenses and the nozzle) as an impactor plate. Particles<br />

with large Stokes numbers will fail to follow streamlines and<br />

will impact on the plate. There<strong>for</strong>e, it is appropriate to use the<br />

spacer Stokes number (Sts) tocharacterize particle impaction<br />

on the orifice plate. The characteristic velocity and length in<br />

Sts are the average flow velocity in the spacer (Us) and the<br />

spacer inner diameter (ds), respectively, i.e., Sts = τUs/ds,<br />

where Us = 4 Q/(π d2 s ). Figures 10(a)–(c) show the particle<br />

transmission efficiency as a function <strong>of</strong> Sts, Re and Ma <strong>for</strong> the<br />

same conditions as those in Figure 2. The corresponding Stokes<br />

numbers based on orifice are also shown in the figures. Note<br />

that significant particle losses happen in the Sts range <strong>of</strong> 0.1–1<br />

<strong>for</strong> all Reynolds and Mach numbers. Although losses in these<br />

figures consist <strong>of</strong> both losses on the orifice plate and the downstream<br />

spacer walls, examination <strong>of</strong> particle trajectories shows<br />

that most losses are due to impaction on the orifice plate <strong>for</strong> these<br />

particular simulations. The transmission curves asymptotically<br />

reach ηt = 2( d f<br />

ds )2 − ( d f<br />

ds )4 <strong>for</strong> very large Stokes numbers corresponding<br />

to geometrical blocking <strong>of</strong> the orifice plate (Zhang<br />

et al. 2002). Figure 11 shows the cut<strong>of</strong>f Stokes number (Sts50)<br />

at which the transmission efficiency is 50% as a function <strong>of</strong> Re<br />

<strong>for</strong> the three Ma’s studied. From Figure 10 we can see that the<br />

impaction particle loss is relatively steep. There<strong>for</strong>e, the particle<br />

transmission efficiency through the stage from lens i–1 to<br />

lens i only considering impaction losses on the orifice plate i,<br />

ηt, orifice, i, can be estimated as a step function:<br />

ηt, orifice, i =<br />

⎧<br />

⎪⎨<br />

1 <strong>for</strong> Sts < Sts50 or dB,i−1 ≤ d f,i<br />

� � d f,i ) 2 � � d 4<br />

f,i<br />

2 d −<br />

s,i−1 d .<br />

s,i−1<br />

⎪⎩<br />

�4 <strong>for</strong> Sts ≥ Sts50 and dB,i−1 > d f,i<br />

� � dB,i−1<br />

2 � dB,i−1<br />

2 d −<br />

s,i−1 ds,i−1 [23]<br />

Note that non-uni<strong>for</strong>m particle concentration due to focusing by<br />

upstream lenses is taken into consideration in this equation.<br />

Defocused particles will be lost to the wall when ηc < −1.<br />

In this case, the initial particle radial position upstream <strong>of</strong> the


FIG. 10. Particle transmission efficiency as a function <strong>of</strong> the Stokes number<br />

<strong>for</strong> various Reynolds numbers at three different subsonic Mach numbers.<br />

(a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.<br />

A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 329<br />

FIG. 11. Cut<strong>of</strong>f Stokes numbers as a function <strong>of</strong> Reynolds number <strong>for</strong> three<br />

different Mach numbers.<br />

lens corresponding to the limiting trajectory <strong>of</strong> particle loss to<br />

the downstream spacer is ri = ds,i/(2ηc,i). The transmission<br />

efficiency can be estimated as the ratio <strong>of</strong> particle flux within ri<br />

to the total flux in that stage. There<strong>for</strong>e,<br />

ηt, spacer, i =<br />

⎧<br />

⎪⎨<br />

1 <strong>for</strong> dB,i ≤ ds,i<br />

⎪⎩<br />

<strong>for</strong> dB,i > ds,i<br />

. [24]<br />

� � d<br />

2 � �<br />

s,i<br />

d<br />

4<br />

s,i<br />

2 ds,i−1η −<br />

c,i ds,i−1ηc,i � � dB,i−1<br />

2 � � dB,i−1<br />

4<br />

2 d −<br />

s,i−1 ds,i−1 Note that dB,i = dB,i−1 ×|ηc,i|, and the diffusion effect is<br />

neglected.<br />

Loss by diffusion at each stage can be estimated through the<br />

Gormley-Kennedy equation (Gormley and Kennedy 1949):<br />

ηt, GK, i =<br />

⎧<br />

⎨1<br />

− 5.50 ξ<br />

⎩<br />

2/3 + 3.77ξ <strong>for</strong> ξ


330 X. WANG AND P. H. MCMURRY<br />

DESCRIPTION OF THE “LENS CALCULATOR”<br />

A Micros<strong>of</strong>t Excel spreadsheet serves as a user interface <strong>for</strong><br />

the <strong>Lens</strong> Calculator. The detailed calculations are done in the<br />

background by a program written in Visual Basic. The calculator<br />

has two modules: lens design and lens test. The lens design<br />

module designs lens dimensions and operating conditions <strong>for</strong><br />

user specified parameters. The lens test module estimates the<br />

FIG. 12. Flow chart <strong>for</strong> the aerodynamic lens design module.<br />

per<strong>for</strong>mance <strong>for</strong> a lens system with specified dimensions and<br />

operating conditions.<br />

Figure 12 shows the flow chart <strong>of</strong> the lens design module. The<br />

design process starts with reading in the user input in<strong>for</strong>mation<br />

<strong>of</strong> carrier gas, several operating parameters, particle density,<br />

and the focusing size range. Then the program calculates the<br />

operating pressure and dimensions <strong>of</strong> the lens system (orifice


A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 331<br />

TABLE 2<br />

Key features <strong>of</strong> the three aerodynamic lens systems used <strong>for</strong> the lens design tool validation<br />

<strong>Lens</strong> system d f (mm) p1 (Pa) dp (nm) QSTP (slm)<br />

1(Wang et al. 2005) 1.26, 1.64, 2.33, 2.76 528 3–30 0.1<br />

2 (Liu et al. 1995) 5.0, 4.5, 4.0, 3.75, 3.5, 3 291 25–250 0.108<br />

3 (Schreiner et al. 2002) 1.4, 1.2, 1.0, 0.8, 0.7, 0.6, 0.5, 0.4 5000 300–5000 0.052<br />

diameters and the length and inner diameter <strong>of</strong> spacers). Finally,<br />

the program estimates the three major per<strong>for</strong>mance parameters:<br />

size dependent particle terminal axial velocity, beam width, and<br />

transmission efficiency.<br />

The user can either specify an operating pressure, or let the<br />

program design the lens system so that it operates at maximum<br />

possible pressure. The latter feature is especially useful when<br />

one tries to minimize the pumping capacity or the particle diffusion<br />

effects. The user can also let the program design lenses<br />

to operate at optimal or user specified Stokes numbers. The<br />

latter feature is typically used when focusing nanoparticles because<br />

sometimes it is not possible to operate lenses at optimal<br />

Stokes numbers while flow is subsonic and continuum (Wang<br />

et al. 2005a, b). It is also useful when designing lenses to operate<br />

at alternating optimal/suboptimal Stokes numbers to focus<br />

a wider size range. The flow through lenses is always <strong>for</strong>ced<br />

to be laminar, continuum, and subsonic. Flow through the pressure<br />

limiting orifice can be turbulent and supersonic, but flow<br />

through the accelerating nozzle is <strong>for</strong>ced to be choked because<br />

otherwise particles may have a short stopping distance and be<br />

lost to the pumps.<br />

The lens test module is very similar to the design module<br />

except that the lens dimensions are specified by the user as well.<br />

The program calculates the operating conditions and estimates<br />

the assembly per<strong>for</strong>mance.<br />

VALIDATION OF THE LENS CALCULATOR<br />

In this section, we compare the per<strong>for</strong>mances (transmission<br />

efficiency and particle beam width) <strong>of</strong> several aerodynamic lens<br />

systems reported in the literature with predictions by the lens<br />

calculator. Three lens systems are selected in this comparison.<br />

<strong>Lens</strong> system 1 was designed by Wang et al. to focus 3–30 nm particles<br />

using helium as the carrier gas (Wang et al. 2005b); <strong>Lens</strong><br />

system 2 was designed by Liu et al. and reported to have good<br />

focusing <strong>for</strong> 25–250 nm particles (lens e in Liu et al. 1995b);<br />

<strong>Lens</strong> system 3 was designed by Schreiner et al. to focus 0.34–4<br />

µm particles in the pressure range <strong>of</strong> 2.5–5 kPa (Schreiner et al.<br />

1999, 2002). These three lens systems cover a wide range <strong>of</strong><br />

focusing sizes and operating pressures. Only numerically simulated<br />

per<strong>for</strong>mance is reported <strong>for</strong> lens system 1, while experimental<br />

data are available <strong>for</strong> lens systems 2 and 3. The major<br />

features <strong>of</strong> these lens systems are listed in Table 2.<br />

Figure 13(a) compares reported particle penetrations through<br />

the above mentioned lenses with results predicted by the lens<br />

calculator. Note that the Calculator predicts particle penetration<br />

reasonably well <strong>for</strong> the three lens systems. For lens system 1,<br />

the prediction is lower than the numerical simulation in the size<br />

range <strong>of</strong> 1.5–40 nm. The first reason is, as we mentioned earlier,<br />

the diffusion loss estimation model in the Calculator is not very<br />

accurate (Equation 26). The second reason is that the particle<br />

penetration through the relaxation chamber was not included in<br />

the numerical simulation (Wang et al. 2005b) but it is included<br />

in the Calculator estimation. The Calculator also over predicts<br />

FIG. 13. Comparison <strong>of</strong> the lens calculator prediction with experimental or<br />

numerical evaluation <strong>for</strong> three lens systems in the literature. (a) Transmission<br />

efficiency; (b) Particle beam width. The legends <strong>of</strong> the two figures are the same.


332 X. WANG AND P. H. MCMURRY<br />

the particle size at which significant impaction losses occurs.<br />

This is most probably due to inaccuracies <strong>of</strong> interpolated contraction<br />

factors from Figure 2 and cut<strong>of</strong>f Stokes numbers from<br />

Figure 11. The discrepancies between the Calculator prediction<br />

and experimental data <strong>of</strong> the lens system are probably due to the<br />

nonideal factors in experiments. The Calculator prediction has<br />

a larger discrepancy with the experimental data <strong>for</strong> lens system<br />

3. Two possibilities contribute to this difference. First, this lens<br />

system operates at larger Reynolds numbers (50–170), where<br />

flow instability might play a role. Second, the spacers between<br />

lenses are very short (15 mm) as compared to suggested values<br />

(∼20–50 mm). There<strong>for</strong>e, recirculation will probably fill the<br />

whole spacer and cause extra losses.<br />

Figure 13(b) compares reported particle beam widths with<br />

values found using the Calculator. We can see that <strong>for</strong> lens system<br />

1, agreement is very good <strong>for</strong> particles smaller than 10 nm.<br />

However, the deviation is significant <strong>for</strong> particles larger than 10<br />

nm. The main reason is that the Calculator cannot predict the<br />

focusing/defocusing behavior <strong>of</strong> the accelerating nozzle very<br />

well. The agreement between the prediction and experiment <strong>for</strong><br />

lens system 2 is quite good. Due to the fact that the design<br />

<strong>of</strong> lens system 3 did not closely follow the guidelines in this<br />

paper (shorter spacer, smaller orifice/spacer diameter ratio), the<br />

agreement between the prediction and measurement <strong>of</strong> beam<br />

width is only fair.<br />

From the above comparisons, we can see that although there<br />

are some discrepancies between the prediction and reported data,<br />

the lens per<strong>for</strong>mance predicted by the Calculator is reasonable<br />

overawide range <strong>of</strong> particle sizes and lens operating conditions.<br />

SUMMARY AND CONCLUSIONS<br />

A s<strong>of</strong>tware package that is able to design and evaluate aerodynamic<br />

lens systems is reported in this article. This lens calculator<br />

has a design module and a test module. The design module<br />

calculates lens dimensions based on the user’s inputs, and the<br />

test module evaluates the focusing per<strong>for</strong>mance <strong>for</strong> a lens system<br />

with specified dimensions and operating conditions. Empirical<br />

relations derived from numerical simulations or experiments<br />

were used in the lens design and per<strong>for</strong>mance estimation.<br />

There<strong>for</strong>e the Calculator provides quick results with reasonable<br />

accuracy.<br />

This article provided details <strong>of</strong> relationships used by the Calculator.<br />

These include the design <strong>of</strong> major components <strong>of</strong> the<br />

lens system: focusing lenses, spacers, the flow limiting orifice,<br />

the relaxation chamber, and the accelerating nozzle. We also<br />

showed graphically how to design the operating pressure <strong>of</strong> a<br />

lens system. Since a generalized expression <strong>of</strong> Stokes number<br />

that is applicable to both continuum and free molecular regimes<br />

is used and the effects <strong>of</strong> diffusion are taken into account, the<br />

lens Calculator applies to lenses <strong>for</strong> nanoparticles or micron<br />

sized particles.<br />

Three per<strong>for</strong>mance parameters <strong>of</strong> a lens system are provided<br />

by the Calculator: particle terminal axial velocity, particle beam<br />

width, and particle transmission efficiency. A <strong>for</strong>mula is used to<br />

correlate the terminal axial velocity to the particle Stokes number<br />

and speed <strong>of</strong> sound <strong>of</strong> the carrier gas. The particle beam diameter<br />

is estimated from contraction factor and the mean square<br />

displacement due to diffusion at each stage. Particle losses due<br />

to inertial impaction and diffusion are accounted <strong>for</strong> when calculating<br />

the transmission efficiency.<br />

We compared the per<strong>for</strong>mance estimations by the lens calculator<br />

with the numerical or experimental results <strong>of</strong> three lens<br />

systems in the literature. The agreement is reasonably good.<br />

NOMENCLATURE<br />

A f<br />

c =<br />

cross sectional area <strong>of</strong> an orifice<br />

√ γ RT1/M speed <strong>of</strong> sound in the carrier gas<br />

Cc<br />

Cunningham slip correction factor<br />

Cd<br />

orifice flow discharge coefficient<br />

D particle diffusion coefficient<br />

dB,i<br />

beam diameter downstream <strong>of</strong> lens i<br />

d f<br />

orifice diameter<br />

dn<br />

nozzle diameter<br />

dp<br />

particle diameter<br />

dp1<br />

maximum particle diameter in the specified<br />

focusing size range<br />

dpN<br />

minimum particle diameter in the specified<br />

focusing size range<br />

dr<br />

minimum diameter <strong>of</strong> the relaxation<br />

chamber<br />

ds<br />

inner diameter <strong>of</strong> the spacer<br />

ds,i<br />

inner diameter <strong>of</strong> the spacer downstream<br />

<strong>of</strong> lens i<br />

dt<br />

step diameter <strong>of</strong> the accelerating nozzle<br />

ER = 1/β = ds/d f orifice expansion ratio<br />

f (vpr) distribution function <strong>of</strong> particle terminal<br />

radial velocity<br />

Kn flow Knudsen number<br />

Kn∗ critical Knudsen number <strong>for</strong> continuum<br />

flow<br />

Knp<br />

particle Knudsen number<br />

h = (ds − d f )/2 orifice step height<br />

L distance from the detector to the nozzle<br />

la<br />

approach length<br />

le<br />

pipe flow entrance length<br />

l f<br />

redevelopment length<br />

lr<br />

reattachment length<br />

ls<br />

length <strong>of</strong> spacers<br />

ls,i<br />

spacer length between lenses i and i + 1<br />

Lr<br />

length <strong>of</strong> the relaxation chamber<br />

Lt<br />

step length <strong>of</strong> the accelerating nozzle<br />

˙m mass flowrate<br />

M molecular weight <strong>of</strong> the carrier gas<br />

Ma Mach number<br />

Ma∗ critical Mach number <strong>for</strong> orifice flow to<br />

be choked


A DESIGN TOOL FOR AERODYNAMIC LENS SYSTEMS 333<br />

p1<br />

pressure upstream <strong>of</strong> an orifice<br />

p2<br />

fully recovered pressure downstream <strong>of</strong><br />

an orifice<br />

pfocusing<br />

pressure upstream <strong>of</strong> a lens <strong>for</strong> focusing<br />

particles <strong>of</strong> a given size<br />

pMa<br />

minimum pressure <strong>for</strong> flow to be subsonic<br />

pmax<br />

maximum operating pressure <strong>of</strong> an aerodynamic<br />

lens<br />

pKn<br />

minimum pressure <strong>for</strong> flow to be continuum<br />

Q volumetric flowrate<br />

Qi<br />

volumetric flowrate at stage i<br />

R universal gas constant<br />

Re flow Reynolds number based on orifice<br />

diameter<br />

Res<br />

flow Reynolds number based on spacer<br />

diameter<br />

ri<br />

critical initial particle radial position<br />

r pi<br />

particle initial radial location in an aerodynamic<br />

lens<br />

S Sutherland constant<br />

Sa<br />

particle axial stopping distance<br />

Sr<br />

particle radial stopping distance<br />

St Stokes number based on orifice diameter<br />

Sto<br />

Sts<br />

Sts50<br />

optimum Stokes number<br />

Stokes number based on spacer diameter<br />

Spacer Stokes number corresponding to<br />

a 50% impaction loss<br />

stage i includes lens i and the downstream<br />

spacer<br />

t<br />

T1<br />

TpF<br />

particle residence time<br />

temperate upstream <strong>of</strong> the lens<br />

particle frozen temperature in the jet expansion<br />

Tr<br />

u<br />

reference temperature in Sutherland’s<br />

Law<br />

average flow velocity at orifice entrance<br />

based on upstream flow conditions<br />

u p<br />

particle axial velocity<br />

Us<br />

va<br />

vpr<br />

average flow velocity in the spacer<br />

axial jet flow velocity<br />

particle terminal radial velocity in the<br />

vacuum chamber<br />

vr<br />

xc ≈ 1 −<br />

radial jet flow velocity<br />

� �p pressure drop across an orifice<br />

ηc<br />

particle contraction factor<br />

ηc,i<br />

contraction factor at lens i<br />

ηt<br />

particle transmission efficiency<br />

ηt, diffusion, i penetration after diffusional loss at stage<br />

i<br />

ηt, GK, i<br />

penetration after loss at stage i estimated<br />

by Gormley-Kennedy equation<br />

ηt, orifice, i<br />

transmission efficiency after impaction<br />

losses on the orifice plate i<br />

ηt, spacer, i<br />

transmission efficiency after impaction<br />

loss to spacer i<br />

θ jet opening angle<br />

λ1<br />

mean free path <strong>of</strong> the gas molecules upstream<br />

<strong>of</strong> the orifice<br />

µ carrier gas viscosity<br />

ξ =<br />

� γ<br />

2 γ −1<br />

γ +1<br />

xrms<br />

particle root mean square displacement<br />

due to diffusion<br />

xrms,i<br />

root mean square displacement between<br />

lenses i and i+1<br />

Y<br />

β = d f /ds<br />

γ<br />

orifice flow expansion factor<br />

constriction ratio<br />

specific heat ratio <strong>of</strong> the carrier gas<br />

Dils,i<br />

Qi<br />

dimensionless diffusion deposition parameter<br />

ρ1<br />

carrier gas density upstream <strong>of</strong> the aerodynamic<br />

lens<br />

ρp<br />

particle material density<br />

τ particle relaxation time<br />

REFERENCES<br />

Back, L. H., and Roschke, E. J. (1972). Shear-Layer Flow Regimes and Wave<br />

Instabilities and Reattachment Lengths Downstream <strong>of</strong> an Abrupt Circular<br />

Channel Expansion, J. App. Mech. 94:677–681.<br />

Bean, H. S. (1971). Fluid Meters: Their Theory and Applications (Report <strong>of</strong><br />

ASME research committee on fluid meters). New York, ASME.<br />

Cheng, Y. S., and Dahneke, B. E. (1979). Properties <strong>of</strong> Continuum Source Particle<br />

Beam. II. Beams Generated in Capillary Expansions, J. Aerosol Sci.<br />

10:363–368.<br />

Dahneke, B. E., and Cheng, Y. S. (1979). Properties <strong>of</strong> Continuum Source<br />

Particle Beam. I. Calculation Methods and Results, J. Aerosol Sci. 10:257–<br />

274.<br />

Di Fonzo, F., Gidwani, A., Fan, M. H., Neumann, A., Iordanoglou, D. I.,<br />

Heberlein, J. V. R., McMurry, P. H., Girshick, S. L., Tymiak, N., Gerberich,<br />

W. W., and Rao, N. P. (2000). Focused Nanoparticle-Beam Deposition <strong>of</strong><br />

Patterned Microstructures, Appl. Phys. Lett. 77(6):910–912.<br />

Dong, Y., Bapat, A., Hilchie, S., Kortshagen, U., and Campbell, S. A. (2004).<br />

Generation <strong>of</strong> Nano-Sized Free Standing Single Crystal Silicon Particles,<br />

J. Vacuum Sci. & Technol. B: Microelectronics and Nanometer Structures<br />

22(4):1923–1930.<br />

Drewnick, F., Hings, S. S., DeCarlo, P., Jayne, J. T., Gonin, M., Fuhrer, K.,<br />

Weimer, S., Jimenez, J. L., Demerjian, K. L., Borrmann, S., and Worsnop, D.<br />

R. (2005). A New Time-<strong>of</strong>-Flight Aerosol Mass Spectrometer(TOF-AMS)—<br />

Instrument Description and First Field Deployment, Aerosol Sci. Technol.<br />

39(7):637–658.<br />

Eichler, T., de Juan, L., and Fernández de la Mora, J. (1998). Improvement <strong>of</strong><br />

the Resolution <strong>of</strong> TSI’s 3071 DMA via Redesigned Sheath Air and Aerosol<br />

Inlets, Aerosol Sci. Technol. 29(1):39–49.<br />

Fernández de la Mora, J., and Riesco-Chueca, P. (1988). <strong>Aerodynamic</strong> Focusing<br />

<strong>of</strong> Particles in a Carrier Gas, J. Fluid Mech. 195:1–21.<br />

Fernández de la Mora, J., Rosell-Llompart, J., and Riesco-Chueca, P. (1989).<br />

<strong>Aerodynamic</strong> Focusing <strong>of</strong> Particles and Molecules in Seeded Supersonic Jets.<br />

in Rarefied Gas Dynamics: Physical Phenomena, Progress in Astronautics &<br />

Aeronautics. E. P. Muntz, D. P. Weaver, and D. H. Campbell, eds., Washington,<br />

DC, AIAA. 117:247–277.


334 X. WANG AND P. H. MCMURRY<br />

Gidwani, A. (2003). Studies <strong>of</strong> Flow and Particle Transport in Hypersonic<br />

Plasma Particle Deposition and <strong>Aerodynamic</strong> Focusing. Ph.D. Thesis,<br />

<strong>Department</strong> <strong>of</strong> Mechanical Engineering. Minneapolis, 55455, University <strong>of</strong><br />

Minnesota.<br />

Girshick, S. L., Heberlein, J. V. R., McMurry, P. H., Gerberich, W. W., Iordanoglou,<br />

D. I., Rao, N. P., Gidwani, A., Tymiak, N., Fonzo, F. D., Fan,<br />

M. H., and Neumann, D. (2000). Hypersonic Plasma Particle Deposition <strong>of</strong><br />

Nanocrystalline Coatings. in Innovative Processing <strong>of</strong> Films and Nanocrystalline<br />

Powders. Choy, K.-L. ed., London, Imperial College Press.<br />

Gómez-Moreno, F. J., Rosell-Llompart, J., and Fernández de la Mora, J. (2002).<br />

Turbulent Transition in Impactor Jets and its Effects on Impactor Resolution,<br />

J. Aerosol Sci. 33:459–476.<br />

Gong, S. C., Liu, R. G., Chou, F. C., and Chiang, A. S. T. (1996). Experiment<br />

and Simulation <strong>of</strong> the Recirculation Flow in a CVD Reactor <strong>for</strong> Monolithic<br />

Materials, Experimental Thermal and Fluid Science 12(1):45–51.<br />

Gormley, P. G., and Kennedy, M. (1949). Diffusion from a Stream Flowing<br />

through a Cylindrical Tube, Proc. Royal Irish Acad. 52(A):163–169.<br />

Huffman, J. A., Jayne, J. T., Drewnick, F., Aiken, A. C., Onasch, T., Worsnop,<br />

D. R., and Jimenez, J. L. (2005). <strong>Design</strong>, Modeling, Optimization, and Experimental<br />

Tests <strong>of</strong> a Particle Beam Width Probe <strong>for</strong> the Aerodyne Aerosol<br />

Mass Spectrometer, Aerosol Sci. Technol. 39(12):1143–1163.<br />

Israel, G. W., and Friedlander, S. K. (1967). High-Speed Beams <strong>of</strong> Small Particles,<br />

J. Coll. Interface Sci. 24:330–337.<br />

Jayne, J. T., Leard, D. C., Zhang, X., Davidovits, P., Smith, K. A., Kolb, C. E., and<br />

Worsnop, D. R. (2000). Development <strong>of</strong> an Aerosol Mass Spectrometer <strong>for</strong><br />

Size and Composition Analysis <strong>of</strong> Submicron Particles, Aerosol Sci. Technol.<br />

33:49–70.<br />

Lee, J.-W., Yi, M.-Y., and Lee, S.-M. (2003). Inertial Focusing <strong>of</strong> Particles with<br />

an <strong>Aerodynamic</strong> <strong>Lens</strong> in the Atmospheric Pressure Range, J. Aerosol Sci.<br />

34:211–234.<br />

Liu, P., Ziemann, P. J., Kittelson, D. B., and McMurry, P. H. (1995a). Generating<br />

Particle Beams <strong>of</strong> Controlled Dimensions and Divergence: I. Theory <strong>of</strong><br />

Particle Motion in <strong>Aerodynamic</strong> <strong>Lens</strong>es and Nozzle Expansions, Aerosol Sci.<br />

Technol. 22(3):293–313.<br />

Liu, P., Ziemann, P. J., Kittelson, D. B., and McMurry, P. H. (1995b). Generating<br />

Particle Beams <strong>of</strong> Controlled Dimensions and Divergence: II. Experimental<br />

Evaluation <strong>of</strong> Particle Motion in <strong>Aerodynamic</strong> <strong>Lens</strong>es and Nozzle Expansions,<br />

Aerosol Sci. Technol. 22(3):314–324.<br />

Mallina, R. V., Wexler, A. S., and Johnston, M. V. (1997). Particle Growth in<br />

High-Speed Particle Beam Inlets, J. Aerosol Sci. 28(2):223–238.<br />

Mallina, R. V., Wexler, A. S., Rhoads, K. P., and Johnston, M. V. (2000). High<br />

Speed Particle Beam Generation: a Dynamic Focusing Mechanism <strong>for</strong> Selecting<br />

Ultrafine Particles, Aerosol Sci. Technol. 33:87–104.<br />

Murphy, W. K., and Sears, G. W. (1964). Production <strong>of</strong> Particulate Beams, J.<br />

Appl. Phys. 35:1986–1987.<br />

Öktem, B., Tolocka, M. P., and Johnston, M. V. (2004). On-Line Analysis <strong>of</strong> Organic<br />

Components in Fine and Ultrafine Particles by Photoionization Aerosol<br />

Mass Spectrometry, Anal. Chem. 76(2):253–261.<br />

Oliveira, P. J., Pinho, F. T., and Schulte, A. (1998). A General Correlation<br />

<strong>for</strong> the Local Loss Coefficient in Newtonian Axisymmetric<br />

Sudden Expansions. International J. Heat and Fluid Flow 19(6):655–<br />

660.<br />

Piseri, P., Tafreshi, H. V., and Milani, P. (2004). Manipulation <strong>of</strong> Nanoparticles<br />

in Supersonic Beams <strong>for</strong> the Production <strong>of</strong> Nanostructured Materials, Curr.<br />

Opini. Solid State Mater. Sci. 8(3–4):195–202.<br />

Schreiner, J., Voigt, C., Mauersburger, K., McMurry, P. H., and Ziemann, P.<br />

(1998). <strong>Aerodynamic</strong> <strong>Lens</strong> System <strong>for</strong> Producing Particle Beams at Stratospheric<br />

Pressures, Aerosol Sci. Technol. 29(1):50–56.<br />

Schreiner, J., Schmid, U., Voigt, C., and Mauersberger, K. (1999). Focusing <strong>of</strong><br />

Aerosols into a Particle Beam at Pressures from 10 to 150 torr, Aerosol Sci.<br />

Technol. 31:373–382.<br />

Schreiner, J., Voigt, C., Zink, P., Kohlmann, A., Knopf, D., Weisser, C., Budz,<br />

P., and Mauersberger, K. (2002). A Mass Spectrometer System <strong>for</strong> Analysis<br />

<strong>of</strong> Polar Stratospheric Aerosols, Rev. Scientif. Instruments 73(2):446–452.<br />

Su, Y., Sipin, M. F., Furutani, H., and Prather, K. A. (2004). Development and<br />

Characterization <strong>of</strong> an Aerosol Time-<strong>of</strong>-Flight Mass Spectrometer with Increased<br />

Detection Efficiency, Anal. Chem. 76(3):712–719.<br />

Svane, M., Hagström, M., and Pettersson, J. B. C. (2004). Chemical Analysis <strong>of</strong><br />

Individual Alkali-Containing Aerosol Particles: <strong>Design</strong> and Per<strong>for</strong>mance <strong>of</strong><br />

a Surface Ionization Particle Beam Mass Spectrometer, Aerosol Sci. Technol.<br />

38(7):655–663.<br />

Tafreshi, H. V., Benedek, G., Piseri, P., Vinati, S., Barborini, E., and Milani,<br />

P. (2002a). A Simple Nozzle Configuration <strong>for</strong> the Production <strong>of</strong> Low Divergence<br />

Supersonic Cluster Beam by <strong>Aerodynamic</strong> Focusing, Aerosol Sci.<br />

Technol. 36:593–606.<br />

Tobias, H. J., Kooiman, P. M., Docherty, K. S., and Ziemann, P. J. (2000). Realtime<br />

Chemical Analysis <strong>of</strong> Organic Aerosol Using a Thermal Desorption<br />

Particle Beam Mass Spectrometer, Aerosol Sci. Technol. 33:170–190.<br />

Wang, X., Kruis, F. E., and McMurry, P. H. (2005a). <strong>Aerodynamic</strong> Focusing <strong>of</strong><br />

Nanoparticles: I. Guidelines <strong>for</strong> <strong>Design</strong>ing <strong>Aerodynamic</strong> <strong>Lens</strong>es <strong>for</strong> Nanoparticles,<br />

Aerosol Sci. Technol. 39(7):611–623.<br />

Wang, X., Gidwani, A., Girshick, S. L., and McMurry, P. H. (2005b). <strong>Aerodynamic</strong><br />

Focusing <strong>of</strong> Nanoparticles: II. Numerical Simulation <strong>of</strong> Particle Motion<br />

through <strong>Aerodynamic</strong> <strong>Lens</strong>es, Aerosol Sci. Technol. 39(7):624–636.<br />

Young, D. F., Munson, B. R., and Okiishi, T. H. (2000), A Brief Introduction to<br />

Fluid Mechanics. John Wiley & Sons, Inc.<br />

Zelenyuk, A., and Imre, D. (2005). Single Particle Laser Ablation Time-<strong>of</strong>-<br />

Flight Mass Spectrometer: An Introduction to SPLAT, Aerosol Sci. Technol.<br />

39(6):554–568.<br />

Zhang, X., Smith, K. A., Worsnop, D. R., Jimenez, J., Jayne, J. T., and Kolb,<br />

C. E. (2002). A Numerical Characterization <strong>of</strong> Particle Beam Collimation by<br />

an <strong>Aerodynamic</strong> <strong>Lens</strong>-Nozzle System: Part I. An Individual <strong>Lens</strong> or Nozzle,<br />

Aerosol Sci. Technol. 36:617–631.<br />

Zhang, X., Smith, K. A., Worsnop, D. R., Jimenez, J. L., Jayne, J. T., Kolb, C. E.,<br />

Morris, J., and Davidovits, P. (2004). Numerical Characterization <strong>of</strong> Particle<br />

Beam Collimation: Part II. Integrated <strong>Aerodynamic</strong>-<strong>Lens</strong>-Nozzle System,<br />

Aerosol Sci. Technol. 38(6):619–638.<br />

Ziemann, P. J., Liu, P., Kittelson, D. B., and McMurry, P. H. (1995). Particle<br />

Beam Mass Spectrometry <strong>of</strong> Submicron Particles Charged to Saturation in an<br />

Electron Beam, J. Aerosol Sci. 26(5):745–756.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!