3,354
Views
85
CrossRef citations to date
0
Altmetric
Original Articles

A Design Tool for Aerodynamic Lens Systems

&
Pages 320-334 | Received 23 Nov 2005, Accepted 03 Feb 2006, Published online: 23 Feb 2007

We report in this article the development of a tool to design and evaluate aerodynamic lens systems: the Aerodynamic Lens Calculator. This Calculator enables quick and convenient design of aerodynamic lens systems based on the parameterized knowledge gained from our detailed numerical simulations of flow and particle transport through aerodynamic lens systems. It designs the key dimensions of a lens system: pressure limiting orifice, relaxation chamber, focusing lenses, spacers and the accelerating nozzle. It also provides estimates of particle terminal axial velocities, particle beam width, and particle transmission efficiencies. This article describes in detail the information that is used in the design tool, including equations for pressure drop through orifices, contraction factors as a function of Reynolds, Mach and Stokes numbers, spacer length as a function of Reynolds number and estimations for the relaxation chamber dimensions. The article also evaluates the performance of the design tool by comparing its predictions with the performance of aerodynamic lens systems that have been described in the literature.

[Supplementary materials are available for this article. Go to the publisher's online edition of Aerosol Science and Technology for the following free supplemental resources: An instruction manual for the Aerodynamic Lens Calculator.]

NOMENCLATURE

A f =

cross sectional area of an orifice

c = =

speed of sound in the carrier gas

C c =

Cunningham slip correction factor

C d =

orifice flow discharge coefficient

D =

particle diffusion coefficient

d B,i =

beam diameter downstream of lens i

d f =

orifice diameter

d n =

nozzle diameter

d p =

particle diameter

d p1 =

maximum particle diameter in the specified focusing size range

d pN =

minimum particle diameter in the specified focusing size range

d r =

minimum diameter of the relaxation chamber

d s =

inner diameter of the spacer

d s,i =

inner diameter of the spacer downstream of lens i

d t =

step diameter of the accelerating nozzle

ER = 1/β = d s /d f =

orifice expansion ratio

f(v pr )=

distribution function of particle terminal radial velocity

Kn =

flow Knudsen number

Kn*=

critical Knudsen number for continuum flow

Kn p =

particle Knudsen number

h = (d s d f )/2=

orifice step height

L=

distance from the detector to the nozzle

l a =

approach length

l e =

pipe flow entrance length

l f =

redevelopment length

l r =

reattachment length

l s =

length of spacers

l s,i =

spacer length between lenses i and i + 1

L r =

length of the relaxation chamber

L t =

step length of the accelerating nozzle

[mdot]=

mass flowrate

M =

molecular weight of the carrier gas

Ma =

Mach number

Ma*=

critical Mach number for orifice flow to be choked

p 1 =

pressure upstream of an orifice

p 2 =

fully recovered pressure downstream of an orifice

p focusing =

pressure upstream of a lens for focusing particles of a given size

p Ma =

minimum pressure for flow to be subsonic

p max =

maximum operating pressure of an aerodynamic lens

p Kn =

minimum pressure for flow to be continuum

Q =

volumetric flowrate

Q i =

volumetric flowrate at stage i

R =

universal gas constant

Re =

flow Reynolds number based on orifice diameter

Re s =

flow Reynolds number based on spacer diameter

r i =

critical initial particle radial position

r pi =

particle initial radial location in an aerodynamic lens

S =

Sutherland constant

S a =

particle axial stopping distance

S r =

particle radial stopping distance

St =

Stokes number based on orifice diameter

St o =

optimum Stokes number

St s =

Stokes number based on spacer diameter

Sts50 =

Spacer Stokes number corresponding to a 50% impaction loss

stage i=

includes lens i and the downstream spacer

t =

particle residence time

T 1 =

temperate upstream of the lens

T pF =

particle frozen temperature in the jet expansion

T r =

reference temperature in Sutherland's Law

u =

average flow velocity at orifice entrance based on upstream flow conditions

u p =

particle axial velocity

U s =

average flow velocity in the spacer

v a =

axial jet flow velocity

v pr =

particle terminal radial velocity in the vacuum chamber

v r =

radial jet flow velocity

x c =

≈ 1 −

x rms =

particle root mean square displacement due to diffusion

x rms, i =

root mean square displacement between lenses i and i+1

Y =

orifice flow expansion factor

β = d f /d s =

constriction ratio

γ=

specific heat ratio of the carrier gas

Δ p =

pressure drop across an orifice

η c =

particle contraction factor

η c,i =

contraction factor at lens i

η t =

particle transmission efficiency

η t, diffusion, i =

penetration after diffusional loss at stage i

η t, GK, i =

penetration after loss at stage i estimated by Gormley-Kennedy equation

η t, orifice, i =

transmission efficiency after impaction losses on the orifice plate i

η t, spacer, i =

transmission efficiency after impaction loss to spacer i

θ=

jet opening angle

λ1 =

mean free path of the gas molecules upstream of the orifice

μ=

carrier gas viscosity

ξ = =

dimensionless diffusion deposition parameter

ρ1 =

carrier gas density upstream of the aerodynamic lens

ρ p =

particle material density

τ=

particle relaxation time

INTRODUCTION

Aerodynamic lens systems have been widely used to generate collimated narrow particle beams since they were invented by CitationLiu et al. (1995a), Citation(b). Typical applications include inlets for aerosol mass spectrometers (CitationZiemann et al. 1995; CitationSchreiner et al. 1998, Citation1999, Citation2002; CitationJayne et al. 2000; CitationTobias et al. 2000; CitationÖktem et al. 2004; CitationSu et al. 2004; CitationSvane et al. 2004; CitationDrewnick et al. 2005; CitationZelenyuk and Imre 2005), nanostructured material synthesis (CitationGirshick et al. 2000; CitationDong et al. 2004; CitationPiseri et al. 2004), and micro-scale device fabrication (CitationDi Fonzo et al. 2000; CitationGidwani 2003).

In the original design of the aerodynamic lens system (CitationLiu et al. 1995a, Citation(b), aerosols are sampled from one atmosphere pressure through a pressure limiting orifice into the lenses which operate at a pressure of ∼200 Pa, with low pressure drop across each lens. Typically a series of three to five lenses progressively focus spherical particles in the size range of 25–250 nm (near unit density) into a narrow beam. The collimated particle beam is then delivered through an accelerating nozzle to the detection chamber maintained at very low pressure. Most lens systems being used today are similar in design to the system described by CitationLiu et al. (1995a), Citation(b).

Several studies have reported on lens systems that were designed to operate under conditions that are significantly different from those used by CitationLiu et al. (1995a), Citation(b). For example, CitationDi Fonzo et al. (2000) designed a lens system to focus silicon carbide particles in the diameter range of 10–100 nm using an argon-hydrogen mixture; CitationDong et al. (2004) used a lens system to focus 10–200 nm silicon particles in hydrogen. Schreiner et al. developed aerodynamic lens systems operating at elevated pressures (2.5–20 kPa) for stratospheric aerosol studies (CitationSchreiner et al. 1998, Citation1999, Citation2002). CitationLee et al. (2003) used a single lens to focus micron-sized particles in the atmospheric pressure range. CitationWang et al. (2005a), Citation(b) developed a systematic procedure to design aerodynamic lenses for 3–30 nm nanoparticles.

To characterize the performance of an aerodynamic lens system, detailed numerical or experimental studies are required. CitationLiu et al. (1995a), Citation(b) carried out theoretical, numerical, and experimental analysis of the performance of aerodynamic lens systems. CitationZhang et al. (2002), Citation(2004) repeated the simulations by Liu et al. with a compressible flow model. To study particle loss and beam broadening due to diffusion, CitationGidwani (2003) and CitationWang et al. (2005b) simulated the flow and particle transport in the lens system considering the particle Brownian motion.

Although numerical simulations or experimental evaluation of the performance of lens systems can provide more accurate results, they are very costly in both time and financial resources. Furthermore, these studies require predefined dimensions and operating conditions of the aerodynamic lens system, which need to be iteratively optimized. Due to the wide applications of aerodynamic lenses, a simple tool that can provide quick and reasonably accurate lens design and evaluation would be useful. We report the development of such an Aerodynamic Lens Calculator in this article. We studied flow and particle transport through a single lens at a wide range of Reynolds and Mach numbers. From these simulations, we obtained information about particle contraction factors, transport efficiencies, flow discharge coefficients through orifices and the lengths of spacer for flow to redevelop. We also simulated flow through the pressure limiting orifice and the relaxation region, the accelerating nozzle, as well as the entire lens system (CitationWang 2005b). Results from these detailed numerical calculations were parameterized and incorporated into the design tool described in this paper. We correlated the particle terminal axial velocity in the vacuum chamber downstream of the nozzle as a function of Stokes number with a simple equation. This enabled us to estimate the particle terminal axial velocity under similar nozzle geometry and operating pressure. We also provide estimate of the particle beam width both inside and downstream of the lens system. Both aerodynamic focusing and diffusion broadening are taken into account in the beam width estimation. We considered three particle loss mechanisms to calculate the particle transport efficiency: inertia impaction on the orifice plate, impaction on the spacer walls due to defocusing, and losses due to diffusion.

This design tool avoids the need for the time consuming detailed numerical simulations, and enables designers to quickly test a range of design options. The tool calculates the lens dimensions, operating conditions and flow parameters at each lens stage. It also estimates the terminal particle velocity after the choked accelerating nozzle, particle beam diameter at each stage and at the location of the detector, and particle transport efficiency through the lens system. This article describes the detailed information used in the Calculator. We first describe the analytical, empirical or numerical relations used to design the dimensions and operating conditions of aerodynamic lens systems. Next we describe the method used to estimate the particle terminal velocity, beam width and transport efficiency. Finally, we report the validation of this lens design tool by comparing its predictions with several detailed numerical and experimental studies reported in the literature.

AERODYNAMIC LENS SYSTEM DESIGN

The objective of aerodynamic lens design is to build a lens system that has best performance, i.e., maximum particle focusing, minimum particle losses, and minimum pumping requirements under user specified inputs (particle size range, particle density, etc.). To achieve this goal, one needs to optimize dimensions of the lens system (number of lenses, lens diameters, inner diameter and length of the spacers between two lenses, the flow limiting orifice, relaxation chamber and the accelerating nozzle) and operating parameters (flowrate, pressure, and carrier gas).

We have briefly described the aerodynamic lens design procedure in an earlier article with emphasis on lenses for nanoparticles (CitationWang et al. 2005a). In this section, we describe the design principle in more detail, and generalize the guidelines to be applicable to particles ranging in size from several nanometers to greater than one micrometer.

Assumptions

shows a schematic of an aerodynamic lens system with the nomenclature used in this paper and in the code of the lens calculator. The lens system described in this article consists of the following parts: an inlet orifice followed by a relaxation chamber, one or more lenses separated by spacers, and an accelerating nozzle. The pressure limiting orifice determines the volumetric flowrate through the lens system, and the accelerating nozzle defines the operating pressure. We assume that the lenses are primarily responsible for all focusing. The focusing or defocusing effects that might be caused by the inlet orifice or the accelerating nozzle are not controlled by the design tool, but the user can optimize these effects by varying the operating pressure. The inner diameters of each spacer are usually the same so as to simplify machining.

FIG. 1 Schematic and nomenclatures of the lens system.

FIG. 1 Schematic and nomenclatures of the lens system.

Particles are assumed to be spherical and electrically neutral. The effects of particle shape have been discussed by CitationLiu et al. (1995a), Citation(b) and CitationHuffman et al. (2005). The design tool restricts flow through the lenses to be laminar, subsonic, and continuum.

It is well known that flow instability and turbulence can disperse particles and destroy focusing. However, the Reynolds number above which these effects contribute to defocusing is not known with certainty. CitationEichler et al. (1998) and CitationGómez-Moreno et al. (2002) reported that turbulent transition in an orifice flow occurred at Reynolds number around 70. CitationBack and Roschke (1972) and CitationGong et al. (1996) found flow instabilities around a Reynolds number of 200 downstream of sudden expansions. Our design tool constrains Reynolds numbers (Re) to be below 200. Therefore,

Although sonic or supersonic nozzles have been widely used to focus particles (CitationMurphy and Sears 1964; CitationIsrael and Friedlander 1967; CitationCheng and Dahneke 1979; CitationDahneke and Cheng 1979; CitationFernández de la Mora et al. 1989; CitationMallina et al. 2000; CitationTafreshi et al. 2002), we constrain the flow through lenses to be subsonic to avoid the complicated influence of shock waves on particle focusing. Therefore, the lens Mach number (Ma) is limited to be smaller than the value corresponding to choked flow (Ma*) (CitationWang et al. 2005a), i.e.,

If the flow is in the free molecular regime, the particle inertia would greatly exceed the drag force and no focusing could be achieved. Furthermore, rarefied gas dynamics is too complicated to handle using our simple design tool. Therefore, we restrict the flow Knudsen number (Kn) to be smaller than 0.1 to assume that the flow is continuum,

Focusing Lens

The parameter that governs particle focusing through an aerodynamic lens is the Stokes number (St), which is defined as the ratio of the particle stopping distance at the average orifice velocity (u) to the orifice diameter (d f ):

where C c = 1 + Kn p (1.257 + 0.4e and Kn p = 2λ1/d p . As shown in earlier studies, the particle contraction factor (η c ), which is the ratio of terminal to initial radial positions of the particles travelling through the lens, is a strong function of the Stokes number: |η c | ≈ 1 when St ≪ 1, | η c | < 1 when St ≈ 1, and | η c | > 1 when St ≫ 1. In other words, small particles follow gas streamlines and thus are not focused, intermediate-sized particles are focused and large particles are defocused by the lens. There exists an optimum Stokes number (St o ) for which η c = 0 (CitationLiu et al. 1995a). Liu et al. showed that the contraction factors are independent of the particle initial radial position for particles that are not too far away from the axis. We refer to these near-axis contraction factors unless stated otherwise.

By rearranging Equation (Equation4), we can calculate the diameter of the lens aperture for optimally focused particles:

Note that the optimal Stokes number is a function of flow Reynolds number, Mach number, and lens geometry (CitationLiu et al. 1995a; CitationZhang et al. 2002). We use thin plate orifices as the focusing elements in this paper. CitationLiu et al. (1995a) showed that the dependence of St o on the constriction ratio (β) of thin plate orifices becomes weak when β ≤ 0.25. To simplify the design process, we use β ≤ 0.25 in the design tool. In principle, one could use β > 0.25 if the η c vs. St curves were known.

To obtain the relation between St o , Ma, and Re, we have carried out numerical simulations of particle motion through a single lens at a range of Mach number (Ma = 0.03–0.32) and Reynolds numbers (Re = 1–100) which cover the typical operating conditions of lenses. In these simulations, β was set equal to 0.2. Particles were injected at an initial radial location (r pi ) of 0.15d s from the lens axis and Brownian motion was neglected. The relationship between the contraction factor and St, Re, and Ma is shown in (a)–(c). Note that the dependence of contraction on Reynolds number decreases as Mach number increases. The dependence of the contraction factor on the particle initial radial location becomes significant for large Stokes numbers (CitationLiu et al. 1995a). Therefore, the near-axis contraction factors in are not very representative at the higher St end of each curve, and significant impaction losses occur for particles with St > 10. shows the optimum Stokes number St o as a function of Re for three different Ma, which corresponds to the St where η c = 0 at each curve in . The value of St o for a specific set of Re and Ma can be interpolated from this data.

FIG. 2 Near-axis particle contraction factor as a function of the Stokes number for various Reynolds numbers at three different subsonic Mach numbers. (a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.

FIG. 2 Near-axis particle contraction factor as a function of the Stokes number for various Reynolds numbers at three different subsonic Mach numbers. (a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.

FIG. 3 Optimum Stokes number as a function of Reynolds number for three different Mach numbers.

FIG. 3 Optimum Stokes number as a function of Reynolds number for three different Mach numbers.

From Equation (Equation5), we can see that the pressure upstream of the lens is an important parameter in calculating the lens diameter. Therefore, we need to accurately estimate the pressure drop across an orifice. We have developed a model for viscous flow through an orifice

as well as for the discharge coefficient (CitationWang et al. 2005a)
The expansion factor Y is calculated using the expression of (CitationBean 1971):
Using Equations (Equation5, Equation6, Equation7, Equation8) we can calculate the diameter of each lens for a given mass flowrate and a pressure at any stage of the lens assembly.

To study the particle size range that a multiple lens system can focus, we found from Equation (Equation4) that in the free molecular regime St is proportional to d p while in the continuum regime it is proportional to d 2 p . If a lens system focuses particles of a decade size range d p1d pN with d p1 = 10 d pN , then the Stokes number of d p1 at the last lens will be 10St o if it is in the free molecular regime, and 100St o if it is in the continuum regime. If by any non-ideal effects particles of d p1were not focused exactly onto the axis, they will start to defocus at the last lens or even at an earlier stage. Therefore, it is probably a good idea to design a lens system to cover approximately a decade of particle size.

Substantial focusing can be achieved even with sub-optimal Stokes numbers when a sufficient number of lenses are used. There are two strategies to focus a wider range of particle sizes. First, one can add additional lenses to refocus the larger particles before they are defocused and lost. Second, one can use first several lenses to focus the largest particles (d p1) to a given tolerance close to the axis, and then design the following lenses at η c (d p1) = − 1 and focus smaller sizes sub-optimally.

If particles are large enough that diffusion is not significant, and the largest particles do not defocus in the final stages, the more lenses used in an assembly, the better focusing can be achieved. In practice, however, increasing the number of lenses increases the difficulty of alignment and adds size and expense. Effects of diffusion increase as the lens length increases. Furthermore, larger particles are more prone to be defocused in an assembly with more lenses. Typically, we have found that a lens assembly consisting of five lenses is sufficient to focus particles of a decade size range. Three or four lenses were used in our previous study of nanoparticle focusing (CitationWang et al. 2005b).

Operating Pressures

From Equation (Equation4), we obtain the pressure (p focusing ) for focusing particles d p using the ideal gas law

Note that C c is a function of p focusing . The minimum pressure for flow to be subsonic (p Ma) and continuum (p Kn) are respectively given as (CitationWang et al. 2005a)
and
shows examples of p Ma, p Kn, and the required p focusing to focus particles of different sizes as functions of the orifice size in air and helium when the flowrate is 0.1 standard liter per minute (slm). The lens operating pressures are those along the p focusing curves where p focusing is wider than both p Ma and p Kn. Note that larger particles have a larger operating pressure window, and that using a lighter carrier gas can increase the operating pressures for smaller sizes which enables focusing them. Also shown in is a “Re warning line,” which corresponds to Re = 200 above which turbulence is likely to degrade focusing.

FIG. 4 The operating pressures for focusing unit density particles of different sizes as a function of orifice size using (a) air and (b) helium as the carrier gas. The flowrate is 0.1 slm. The two dash lines are the lower pressure limits p Ma and p Kn, respectively. The solid lines are p focusing for indicated sizes.

FIG. 4 The operating pressures for focusing unit density particles of different sizes as a function of orifice size using (a) air and (b) helium as the carrier gas. The flowrate is 0.1 slm. The two dash lines are the lower pressure limits p Ma and p Kn, respectively. The solid lines are p focusing for indicated sizes.

As was pointed out previously (CitationWang et al. 2005a), increasing the operating pressure can reduce the detrimental effects of nanoparticle diffusion and can reduce the pumping needs. From , we can see that the maximum operating pressure of smaller sizes occurs where the p Ma curve intersects the p focusing curve for each particle size, and its value can be derived from Equations (Equation9) and (Equation10),

Note that this is an implicit expression because C c is a function of p max, and an iterative method is needed to calculate p max. The p Ma curve may not intersect the p focusing curve for larger particles in the practical orifice diameter range. In this case any value of p focusing larger than p Kn can be used. However, shows that the higher the operating pressure, the larger is the Reynolds number. It follows that Reynolds numbers tend to limit the maximum pressure at which lenses can be operated.

Length of Spacers

Spacers are used to separate lenses. They provide room for generating the periodic converging/diverging flow pattern with orifices, which drives aerodynamic focusing. In addition to the constraints mentioned earlier (β ≤ 0.25; equal inner diameter of all spacers), the spacers should be long enough so that the flow from the preceding lens can relax back to fully developed pipe flow before reaching the next lens (CitationLiu et al. 1995b). Typically, spacer lengths are 10–50 times the upstream orifice diameter depending on the orifice Reynolds number (CitationWang et al. 2005a). However, no quantitative guideline on the spacer length has been reported.

shows the flow field through two lenses separated by a spacer. The flow separates from the wall when it passes through the lens and a recirculation zone forms downstream of the orifice. After a distance l r (reattachment length), the flow reattaches to the wall, and becomes fully developed at a distance of l f (redevelopment length) downstream of the orifice. The existence of the downstream lens is felt by the flow l a (approach length) upstream of the next lens where the flow starts to curve toward the centerline. Ideally, the minimum length of spacers (l s ) should be l s1 = l r + l a , and more safely, l s2 = l f + l a . If the spacer length is less than l s1, the recirculation zone will likely fill the whole length of the spacer and less focusing will occur. Furthermore, particles are more likely to be entrained in the recirculation zones and be lost to the wall. Therefore, when particle diffusion is negligible and the overall length of the lens system is not a concern, we suggest using spacer lengths at least l s2. When diffusion is important, l s1 < l s l s2 can be used.

FIG. 5 Streamlines through two lenses separated by a spacer.

FIG. 5 Streamlines through two lenses separated by a spacer.

To estimate the values of l r , l f , and l a , we have carried out numerical simulations of flow through single lenses in the Reynolds number range of 0.1–200. Calculated values for l r are shown in , together with data provided by other researchers (CitationBack and Roschke 1972; CitationOliveira et al. 1998). Note that our numerical results are very close to those by CitationOliveria (1998), which apply to similar Reynolds numbers and expansion ratios. The results by Back et al. differ somewhat from ours and those of Oliveria, probably due to the difference in ER. Since we did not carry out simulations for Re > 200, we will use the data by Back et al. to estimate l r in this Re range, noting that applying Back's data (ER = 2.6) to larger ER values might lead to some error. Also shown in is a fitted curve to the whole Reynolds number range, which can be described as follows

FIG. 6 Reattachment length of flow downstream of an axisymmetric expansion.

FIG. 6 Reattachment length of flow downstream of an axisymmetric expansion.
The redevelopment length l f can be estimated by l f = l r + l e where l e is given as (CitationYoung et al. 2000)
The approach length l a is found in our simulation to be independent of Reynolds number (Re < 150) with
Since l a is relatively shorter than l r and l f , we use the above equation to estimate l a for the whole Reynolds number range.

Flow Limiting Orifice, Relaxation Chamber, and Accelerating Nozzle

The flow limiting orifice, relaxation chamber, and the accelerating nozzle are important components of the aerodynamic lens system. The flow limiting orifice and relaxation chamber are required if the particle source pressure is higher than the lens operating pressure. The orifice sets the volumetric flowrate through the lens system, and reduces the pressure to the lens operating pressure. Usually the flow through the flow limiting orifice is critical. Therefore, a relaxation chamber is needed to slow down the high velocity flow and particles to prevent particle impaction on the downstream lens. The relaxation chamber also provides space for the flow to reattach to the wall after the recirculation eddies downstream of the orifice. The accelerating nozzle controls the exact lens operating pressure, and it accelerates particles to a downstream destination with minimized divergence angles.

The inner diameter of the relaxation chamber can be estimated from the radial stopping distance of the largest particles of interest. The lower half of shows the flow streamlines downstream of a straight-bored critical orifice with a diameter and wall thickness of 0.1 mm. The inlet pressure is 1 atm and the outlet pressure is 266 Pa. The flowrate is 70.6 sccm, which correspond to a Reynolds number of 1071 based on the orifice diameter. Therefore, the jet is turbulent. However, for the sake of simplicity, we only used a steady laminar flow model in this simulation and hence the streamlines in only serve as a qualitative description of the averaged flow behavior. The upper half of shows the half jet opening angle θ, which is defined as θ = tan − 1(h/l r ) ≈ tan − 1(v r /v a ). Assuming v a equals the speed of sound c, then v r c tan (θ). We further conservatively assume that particle velocity equals to the flow velocity, then the particle radial stopping distance can be estimated from S r = τ v r where τ is the relaxation time of the largest particles in the target focusing size range downstream of orifice. The minimum diameter of the relaxation chamber (d r ) is then

FIG. 7 Straight bored critical orifice and cylindrical relaxation chamber. (a) Flow streamlines and illustration of the half jet opening angle; (b) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles.

FIG. 7 Straight bored critical orifice and cylindrical relaxation chamber. (a) Flow streamlines and illustration of the half jet opening angle; (b) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles.
Here d f is the diameter of the critical orifice.

The relaxation chamber should be long enough to let flow reattach and particles slow down. The length l f for flow to redevelop can be calculated using Equations (Equation13) and (Equation14). Due to the complicated shock structure, we assume that particles have velocities equal to the speed of sound c for the carrier gas at a distance l r downstream of the orifice while the gas velocity is negligibly low. The particle axial stopping distance downstream of l r is then S a = τ c. Therefore, the length of the relaxation chamber L r can be estimated as

The shape of the relaxation chamber and critical orifice aperture are also important to reduce particle losses. shows trajectories of 100 nm (above the axis) and 1 μ m (below the axis) particles through the critical orifice. The turbulent dispersion of particles is not included in the simulation. Note that some particles of both sizes are trapped inside the recirculation region for a long time, which makes them susceptible to coagulation or deposition losses. Therefore, we recommend that the strong recirculation region be eliminated. This can be achieved by inserting a conical expansion downstream of the orifice as shown in . shows trajectories of 100 nm and 1 μ m particles through this modified relaxation chamber. Although this new design eliminates the recirculation eddy, impaction losses for larger particles increase due to the narrowed flow path, as is illustrated by the 1 μ m trajectories. To reduce impaction losses of larger particles, one can use conical nozzles, short capillaries or their combinations instead of thin plate orifices, since these nozzle shapes will reduce particle accelerations at the nozzle inlet (CitationFernández de la Mora and Riesco-Chueca 1988; CitationFernández de la Mora et al. 1989). illustrates that impaction losses of 1 μ m particles are reduced by adding a 0.1 mm thick 45° chamfer to the original orifice. (Note that this orifice modification increases the flowrate by 10.4%.) Although we demonstrated methods to improve the performance of the relaxation region in this section, we should point out that more careful studies of the orifice shape and chamber dimensions are still required.

FIG. 8 Modified relaxation chamber with a conical divergent section to reduce recirculation and particle loss. (a) Flow streamlines (straight orifice); (b) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles (straight orifice); (c) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles (orifice with a chamfer).

FIG. 8 Modified relaxation chamber with a conical divergent section to reduce recirculation and particle loss. (a) Flow streamlines (straight orifice); (b) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles (straight orifice); (c) Trajectories of 100 nm (above axis) and 1 μ m (below axis) particles (orifice with a chamfer).

We follow the recommendation of CitationLiu et al. (1995b) to use a step nozzle as the accelerating nozzle of lens systems (see ). The diameter of the final orifice in the nozzle can be calculated using Equation (Equation6), and other dimensions of the nozzle are taken as d t ≈ 2d n , L t d s . Again, further research is needed to optimize the nozzle design so that the nozzle can also achieve maximum focusing for a given size range and operating conditions.

LENS PERFORMANCE ESTIMATION

The most important performance parameters for an aerodynamic lens system are: particle terminal velocities, particle beam widths at various locations, and the particle transmission efficiencies. These parameters need to be evaluated by detailed numerical simulations, and ultimately by experiments. However, it is attractive to have a best estimation of the lens performance with a “one-button-click” effort during the design process. In this section, we describe the method used in the Lens Calculator to do the estimations.

Estimation of Particle Terminal Velocities

We assume that the pressure downstream of the accelerating nozzle is low enough so that the gas-particle collisions are negligible and particles achieve their “terminal” velocities. We further assume that particles are in thermal equilibrium with the carrier gas upstream of the nozzle exit and their radial velocity can be approximately described by the Maxwell-Boltzmann distribution. This velocity distribution can be assumed to be frozen during the expansion downstream of the nozzle (CitationLiu et al. 1995a; CitationWang et al. 2005b). Therefore, the terminal radial velocity v pr can be estimated as

The particle terminal axial velocities (u p ) depend on particle size, shape, nozzle geometry, pressure ratio, and carrier gases (CitationCheng and Dahneke 1979; CitationDahneke and Cheng 1979; CitationMallina et al. 1997). summarizes the particle terminal axial velocities for the four aerodynamic lens systems listed in . These four lens systems differ in dimensions, flowrate, pressure, and carrier gas. But they all use step accelerating nozzles similar to that described by CitationLiu et al. (1995b), and the pressures downstream of the nozzles are less than 1 Pa. Note that the following equation fits the data points reasonably well in the St = τ c/d n range where data is available (0.01 < St < 80)

FIG. 9 Normalized particle terminal axial velocities downstream of several aerodynamic lens systems.

FIG. 9 Normalized particle terminal axial velocities downstream of several aerodynamic lens systems.

TABLE 1 Key features of four lens assemblies with step accelerating nozzles used in particle axial velocity comparison

We will use this equation to estimate particle axial velocity in the Lens Calculator.

Estimation of Particle Beam Widths

The particle beam width is defined as the beam diameter that encloses 90% of the total particle flux. In this section expressions are given that are used in the Calculator to calculate particle beam widths downstream of each lens as well as downstream of the accelerating nozzle. Bean widths are controlled by two factors: aerodynamic focusing and diffusion broadening. The focusing can be inferred from , and the diffusion broadening can be estimated by the root mean square displacement x rms = , where D is the particle diffusion coefficient, and t is the particle residence time.

Assuming the diameters of spacers upstream and downstream of the pressure limiting orifice (lens 0) are the same, the flow is fully developed upstream of the pressure limiting orifice, and particles are homogenously distributed in the flow cross section, the starting particle beam diameter enclosing 90% flux is then ∼0.827d s,0. The particle beam diameter before reaching lens 1 is

The beam diameter d B,i at stage i inside the lens system can be easily calculated as
However, we should note that d B, i may exceed the spacer diameter d s,i due to defocusing or diffusion. In that case, we assume all particles outside the beam diameter of d s,i are lost and reset d B,i = 0.827 d s,i .

Assuming particles achieve a Maxwell-Boltzmann distribution of radial velocities downstream of the acceleration nozzle (lens n + 1), we can estimate the particle beam width at a distance L downstream of the nozzle as follows:

Note that η c,n + 1is only a very rough estimate because the pressure downstream of the nozzle is so low that the definition of contraction factor is no longer valid.

Estimation of Particle Transmission Efficiencies

Particle losses in an aerodynamic lens system arise from impaction and diffusion. Particle impaction happens both on the orifice plate and on the spacer walls.

We can view the orifices (including the pressure limiting orifice, lenses and the nozzle) as an impactor plate. Particles with large Stokes numbers will fail to follow streamlines and will impact on the plate. Therefore, it is appropriate to use the spacer Stokes number (St s ) to characterize particle impaction on the orifice plate. The characteristic velocity and length in St s are the average flow velocity in the spacer (U s ) and the spacer inner diameter (d s ), respectively, i.e., St s = τ U s /d s , where U s = 4 Q/(π d 2 s ). , , show the particle transmission efficiency as a function of St s , Re and Ma for the same conditions as those in . The corresponding Stokes numbers based on orifice are also shown in the figures. Note that significant particle losses happen in the St s range of 0.1–1 for all Reynolds and Mach numbers. Although losses in these figures consist of both losses on the orifice plate and the downstream spacer walls, examination of particle trajectories shows that most losses are due to impaction on the orifice plate for these particular simulations. The transmission curves asymptotically reach η t = for very large Stokes numbers corresponding to geometrical blocking of the orifice plate (CitationZhang et al. 2002). shows the cutoff Stokes number (Sts50) at which the transmission efficiency is 50% as a function of Re for the three Ma's studied. From we can see that the impaction particle loss is relatively steep. Therefore, the particle transmission efficiency through the stage from lens i–1 to lens i only considering impaction losses on the orifice plate i, η t, orifice, i , can be estimated as a step function:

FIG. 10 Particle transmission efficiency as a function of the Stokes number for various Reynolds numbers at three different subsonic Mach numbers. (a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.

FIG. 10 Particle transmission efficiency as a function of the Stokes number for various Reynolds numbers at three different subsonic Mach numbers. (a) Ma = 0.03; (b) Ma = 0.10; (c) Ma = 0.32.

FIG. 11 Cutoff Stokes numbers as a function of Reynolds number for three different Mach numbers.

FIG. 11 Cutoff Stokes numbers as a function of Reynolds number for three different Mach numbers.
Note that non-uniform particle concentration due to focusing by upstream lenses is taken into consideration in this equation.

Defocused particles will be lost to the wall when η c < − 1. In this case, the initial particle radial position upstream of the lens corresponding to the limiting trajectory of particle loss to the downstream spacer is r i = d s,i /(2η c,i ). The transmission efficiency can be estimated as the ratio of particle flux within r i to the total flux in that stage. Therefore,

Note that d B,i = d B,i − 1× |η c,i |, and the diffusion effect is neglected.

Loss by diffusion at each stage can be estimated through the Gormley-Kennedy equation (CitationGormley and Kennedy 1949):

Note that we have assumed fully developed laminar flow in the lens system by using this equation. The actual diffusion loss is less than that predicted by the Gormley-Kennedy equation, because lenses move particles away from walls (CitationWang et al. 2005b). Therefore, the actual penetration accounting for diffusion falls somewhere between the Gormley-Kennedy prediction and 1 when particles are focused to a certain degree. We use the following simplified equation to predict diffusional loss:
The total particle transmission efficiency through the lens system is then

DESCRIPTION OF THE “LENS CALCULATOR”

A Microsoft Excel spreadsheet serves as a user interface for the Lens Calculator. The detailed calculations are done in the background by a program written in Visual Basic. The calculator has two modules: lens design and lens test. The lens design module designs lens dimensions and operating conditions for user specified parameters. The lens test module estimates the performance for a lens system with specified dimensions and operating conditions.

shows the flow chart of the lens design module. The design process starts with reading in the user input information of carrier gas, several operating parameters, particle density, and the focusing size range. Then the program calculates the operating pressure and dimensions of the lens system (orifice diameters and the length and inner diameter of spacers). Finally, the program estimates the three major performance parameters: size dependent particle terminal axial velocity, beam width, and transmission efficiency.

FIG. 12 Flow chart for the aerodynamic lens design module.

FIG. 12 Flow chart for the aerodynamic lens design module.

The user can either specify an operating pressure, or let the program design the lens system so that it operates at maximum possible pressure. The latter feature is especially useful when one tries to minimize the pumping capacity or the particle diffusion effects. The user can also let the program design lenses to operate at optimal or user specified Stokes numbers. The latter feature is typically used when focusing nanoparticles because sometimes it is not possible to operate lenses at optimal Stokes numbers while flow is subsonic and continuum (CitationWang et al. 2005a, b). It is also useful when designing lenses to operate at alternating optimal/suboptimal Stokes numbers to focus a wider size range. The flow through lenses is always forced to be laminar, continuum, and subsonic. Flow through the pressure limiting orifice can be turbulent and supersonic, but flow through the accelerating nozzle is forced to be choked because otherwise particles may have a short stopping distance and be lost to the pumps.

The lens test module is very similar to the design module except that the lens dimensions are specified by the user as well. The program calculates the operating conditions and estimates the assembly performance.

VALIDATION OF THE LENS CALCULATOR

In this section, we compare the performances (transmission efficiency and particle beam width) of several aerodynamic lens systems reported in the literature with predictions by the lens calculator. Three lens systems are selected in this comparison. Lens system 1 was designed by Wang et al. to focus 3–30 nm particles using helium as the carrier gas (CitationWang et al. 2005b); Lens system 2 was designed by Liu et al. and reported to have good focusing for 25–250 nm particles (lens e in CitationLiu et al. 1995b); Lens system 3 was designed by Schreiner et al. to focus 0.34–4 μ m particles in the pressure range of 2.5–5 kPa (CitationSchreiner et al. 1999, Citation2002). These three lens systems cover a wide range of focusing sizes and operating pressures. Only numerically simulated performance is reported for lens system 1, while experimental data are available for lens systems 2 and 3. The major features of these lens systems are listed in .

TABLE 2 Key features of the three aerodynamic lens systems used for the lens design tool validation

compares reported particle penetrations through the above mentioned lenses with results predicted by the lens calculator. Note that the Calculator predicts particle penetration reasonably well for the three lens systems. For lens system 1, the prediction is lower than the numerical simulation in the size range of 1.5–40 nm. The first reason is, as we mentioned earlier, the diffusion loss estimation model in the Calculator is not very accurate (Equation Equation26). The second reason is that the particle penetration through the relaxation chamber was not included in the numerical simulation (CitationWang et al. 2005b) but it is included in the Calculator estimation. The Calculator also over predicts the particle size at which significant impaction losses occurs. This is most probably due to inaccuracies of interpolated contraction factors from and cutoff Stokes numbers from . The discrepancies between the Calculator prediction and experimental data of the lens system are probably due to the nonideal factors in experiments. The Calculator prediction has a larger discrepancy with the experimental data for lens system 3. Two possibilities contribute to this difference. First, this lens system operates at larger Reynolds numbers (50–170), where flow instability might play a role. Second, the spacers between lenses are very short (15 mm) as compared to suggested values (∼ 20–50 mm). Therefore, recirculation will probably fill the whole spacer and cause extra losses.

FIG. 13 Comparison of the lens calculator prediction with experimental or numerical evaluation for three lens systems in the literature. (a) Transmission efficiency; (b) Particle beam width. The legends of the two figures are the same.

FIG. 13 Comparison of the lens calculator prediction with experimental or numerical evaluation for three lens systems in the literature. (a) Transmission efficiency; (b) Particle beam width. The legends of the two figures are the same.

compares reported particle beam widths with values found using the Calculator. We can see that for lens system 1, agreement is very good for particles smaller than 10 nm. However, the deviation is significant for particles larger than 10 nm. The main reason is that the Calculator cannot predict the focusing/defocusing behavior of the accelerating nozzle very well. The agreement between the prediction and experiment for lens system 2 is quite good. Due to the fact that the design of lens system 3 did not closely follow the guidelines in this paper (shorter spacer, smaller orifice/spacer diameter ratio), the agreement between the prediction and measurement of beam width is only fair.

From the above comparisons, we can see that although there are some discrepancies between the prediction and reported data, the lens performance predicted by the Calculator is reasonable over a wide range of particle sizes and lens operating conditions.

SUMMARY AND CONCLUSIONS

A software package that is able to design and evaluate aerodynamic lens systems is reported in this article. This lens calculator has a design module and a test module. The design module calculates lens dimensions based on the user's inputs, and the test module evaluates the focusing performance for a lens system with specified dimensions and operating conditions. Empirical relations derived from numerical simulations or experiments were used in the lens design and performance estimation. Therefore the Calculator provides quick results with reasonable accuracy.

This article provided details of relationships used by the Calculator. These include the design of major components of the lens system: focusing lenses, spacers, the flow limiting orifice, the relaxation chamber, and the accelerating nozzle. We also showed graphically how to design the operating pressure of a lens system. Since a generalized expression of Stokes number that is applicable to both continuum and free molecular regimes is used and the effects of diffusion are taken into account, the lens Calculator applies to lenses for nanoparticles or micron sized particles.

Three performance parameters of a lens system are provided by the Calculator: particle terminal axial velocity, particle beam width, and particle transmission efficiency. A formula is used to correlate the terminal axial velocity to the particle Stokes number and speed of sound of the carrier gas. The particle beam diameter is estimated from contraction factor and the mean square displacement due to diffusion at each stage. Particle losses due to inertial impaction and diffusion are accounted for when calculating the transmission efficiency.

We compared the performance estimations by the lens calculator with the numerical or experimental results of three lens systems in the literature. The agreement is reasonably good.

Supplemental excel

Supplemental material

uast_161487_supplementary_files.zip

Download Zip (2.2 MB)

Acknowledgments

This work was supported by NSF (Grant No. DMI-0103169) and the University of Minnesota Supercomputing Institute. We thank Dr. Frank Einar Kruis for helpful discussions.

REFERENCES

  • Back , L. H. and Roschke , E. J. 1972 . Shear-Layer Flow Regimes and Wave Instabilities and Reattachment Lengths Downstream of an Abrupt Circular Channel Expansion . J. App. Mech. , 94 : 677 – 681 . [CSA]
  • Bean , H. S. 1971 . Fluid Meters: Their Theory and Applications (Report of ASME research committee on fluid meters) , New York : ASME .
  • Cheng , Y. S. and Dahneke , B. E. 1979 . Properties of Continuum Source Particle Beam. II. Beams Generated in Capillary Expansions . J. Aerosol Sci. , 10 : 363 – 368 . [CROSSREF] [CSA]
  • Dahneke , B. E. and Cheng , Y. S. 1979 . Properties of Continuum Source Particle Beam. I. Calculation Methods and Results . J. Aerosol Sci. , 10 : 257 – 274 . [CROSSREF] [CSA]
  • Di Fonzo , F. , Gidwani , A. , Fan , M. H. , Neumann , A. , Iordanoglou , D. I. , Heberlein , J. V. R. , McMurry , P. H. , Girshick , S. L. , Tymiak , N. , Gerberich , W. W. and Rao , N. P. 2000 . Focused Nanoparticle-Beam Deposition of Patterned Microstructures . Appl. Phys. Lett. , 77 ( 6 ) : 910 – 912 . [CROSSREF] [CSA]
  • Dong , Y. , Bapat , A. , Hilchie , S. , Kortshagen , U. and Campbell , S. A. 2004 . Generation of Nano-Sized Free Standing Single Crystal Silicon Particles . J. Vacuum Sci. & Technol. B: Microelectronics and Nanometer Structures , 22 ( 4 ) : 1923 – 1930 . [CROSSREF] [CSA]
  • Drewnick , F. , Hings , S. S. , DeCarlo , P. , Jayne , J. T. , Gonin , M. , Fuhrer , K. , Weimer , S. , Jimenez , J. L. , Demerjian , K. L. , Borrmann , S. and Worsnop , D. R. 2005 . A New Time-of-Flight Aerosol Mass Spectrometer(TOF-AMS)—Instrument Description and First Field Deployment . Aerosol Sci. Technol. , 39 ( 7 ) : 637 – 658 . [CROSSREF] [CSA]
  • Eichler , T. , de Juan , L. and Fernández de la Mora , J. 1998 . Improvement of the Resolution of TSI's 3071 DMA via Redesigned Sheath Air and Aerosol Inlets . Aerosol Sci. Technol. , 29 ( 1 ) : 39 – 49 . [CSA]
  • Fernández , d e , la Mora , J. and Riesco-Chueca , P. 1988 . Aerodynamic Focusing of Particles in a Carrier Gas . J. Fluid Mech. , 195 : 1 – 21 . [CSA]
  • Fernández de la Mora , J. , Rosell-Llompart , J. and Riesco-Chueca , P. 1989 . “ Aerodynamic Focusing of Particles and Molecules in Seeded Supersonic Jets ” . In Rarefied Gas Dynamics: Physical Phenomena, Progress in Astronautics & Aeronautics , Edited by: Muntz , E. P. , Weaver , D. P. and Campbell , D. H. Vol. 117 , 247 – 277 . Washington, DC : AIAA .
  • Gidwani , A. 2003 . Studies of Flow and Particle Transport in Hypersonic Plasma Particle Deposition and Aerodynamic Focusing , University of Minnesota . Ph.D. Thesis, Department of Mechanical Engineering. Minneapolis, 55455
  • Girshick , S. L. , Heberlein , J. V. R. , McMurry , P. H. , Gerberich , W. W. , Iordanoglou , D. I. , Rao , N. P. , Gidwani , A. , Tymiak , N. , Fonzo , F. D. , Fan , M. H. and Neumann , D. 2000 . “ Hypersonic Plasma Particle Deposition of Nanocrystalline Coatings ” . In Innovative Processing of Films and Nanocrystalline Powders , Edited by: Choy , K.-L. London : Imperial College Press .
  • Gómez-Moreno , F. J. , Rosell-Llompart , J. and Fernández de la Mora , J. 2002 . Turbulent Transition in Impactor Jets and its Effects on Impactor Resolution . J. Aerosol Sci. , 33 : 459 – 476 . [CROSSREF] [CSA]
  • Gong , S. C. , Liu , R. G. , Chou , F. C. and Chiang , A. S. T. 1996 . Experiment and Simulation of the Recirculation Flow in a CVD Reactor for Monolithic Materials . Experimental Thermal and Fluid Science , 12 ( 1 ) : 45 – 51 . [CROSSREF] [CSA]
  • Gormley , P. G. and Kennedy , M. 1949 . Diffusion from a Stream Flowing through a Cylindrical Tube . Proc. Royal Irish Acad. , 52 ( A ) : 163 – 169 . [CSA]
  • Huffman , J. A. , Jayne , J. T. , Drewnick , F. , Aiken , A. C. , Onasch , T. , Worsnop , D. R. and Jimenez , J. L. 2005 . Design, Modeling, Optimization, and Experimental Tests of a Particle Beam Width Probe for the Aerodyne Aerosol Mass Spectrometer . Aerosol Sci. Technol. , 39 ( 12 ) : 1143 – 1163 . [CROSSREF] [CSA]
  • Israel , G. W. and Friedlander , S. K. 1967b . High-Speed Beams of Small Particles . J. Coll. Interface Sci. , 24 : 330 – 337 . [CROSSREF] [CSA]
  • Jayne , J. T. , Leard , D. C. , Zhang , X. , Davidovits , P. , Smith , K. A. , Kolb , C. E. and Worsnop , D. R. 2000b . Development of an Aerosol Mass Spectrometer for Size and Composition Analysis of Submicron Particles . Aerosol Sci. Technol. , 33 : 49 – 70 . [CROSSREF] [CSA]
  • Lee , J.-W. , Yi , M.-Y. and Lee , S.-M. 2003b . Inertial Focusing of Particles with an Aerodynamic Lens in the Atmospheric Pressure Range . J. Aerosol Sci. , 34 : 211 – 234 . [CROSSREF] [CSA]
  • Liu , P. , Ziemann , P. J. , Kittelson , D. B. and McMurry , P. H. 1995a . Generating Particle Beams of Controlled Dimensions and Divergence: I. Theory of Particle Motion in Aerodynamic Lenses and Nozzle Expansions . Aerosol Sci. Technol. , 22 ( 3 ) : 293 – 313 . [CSA]
  • Liu , P. , Ziemann , P. J. , Kittelson , D. B. and McMurry , P. H. 1995b . Generating Particle Beams of Controlled Dimensions and Divergence: II. Experimental Evaluation of Particle Motion in Aerodynamic Lenses and Nozzle Expansions . Aerosol Sci. Technol. , 22 ( 3 ) : 314 – 324 . [CSA]
  • Mallina , R. V. , Wexler , A. S. and Johnston , M. V. 1997 . Particle Growth in High-Speed Particle Beam Inlets . J. Aerosol Sci. , 28 ( 2 ) : 223 – 238 . [CROSSREF] [CSA]
  • Mallina , R. V. , Wexler , A. S. , Rhoads , K. P. and Johnston , M. V. 2000 . High Speed Particle Beam Generation: a Dynamic Focusing Mechanism for Selecting Ultrafine Particles . Aerosol Sci. Technol. , 33 : 87 – 104 . [CROSSREF] [CSA]
  • Murphy , W. K. and Sears , G. W. 1964b . Production of Particulate Beams . J. Appl. Phys. , 35 : 1986 – 1987 . [CROSSREF] [CSA]
  • Öktem , B. , Tolocka , M. P. and Johnston , M. V. 2004 . On-Line Analysis of Organic Components in Fine and Ultrafine Particles by Photoionization Aerosol Mass Spectrometry . Anal. Chem. , 76 ( 2 ) : 253 – 261 . [CROSSREF] [CSA]
  • Oliveira , P. J. , Pinho , F. T. and Schulte , A. 1998 . A General Correlation for the Local Loss Coefficient in Newtonian Axisymmetric Sudden Expansions . International J. Heat and Fluid Flow , 19 ( 6 ) : 655 – 660 . [CSA]
  • Piseri , P. , Tafreshi , H. V. and Milani , P. 2004 . Manipulation of Nanoparticles in Supersonic Beams for the Production of Nanostructured Materials . Curr. Opini. Solid State Mater. Sci. , 8 ( 3–4 ) : 195 – 202 . [CROSSREF] [CSA]
  • Schreiner , J. , Voigt , C. , Mauersburger , K. , McMurry , P. H. and Ziemann , P. 1998 . Aerodynamic Lens System for Producing Particle Beams at Stratospheric Pressures . Aerosol Sci. Technol. , 29 ( 1 ) : 50 – 56 . [CSA]
  • Schreiner , J. , Schmid , U. , Voigt , C. and Mauersberger , K. 1999b . Focusing of Aerosols into a Particle Beam at Pressures from 10 to 150 torr . Aerosol Sci. Technol. , 31 : 373 – 382 . [CROSSREF] [CSA]
  • Schreiner , J. , Voigt , C. , Zink , P. , Kohlmann , A. , Knopf , D. , Weisser , C. , Budz , P. and Mauersberger , K. 2002 . A Mass Spectrometer System for Analysis of Polar Stratospheric Aerosols . Rev. Scientif. Instruments , 73 ( 2 ) : 446 – 452 . [CROSSREF] [CSA]
  • Su , Y. , Sipin , M. F. , Furutani , H. and Prather , K. A. 2004 . Development and Characterization of an Aerosol Time-of-Flight Mass Spectrometer with Increased Detection Efficiency . Anal. Chem. , 76 ( 3 ) : 712 – 719 . [PUBMED] [INFOTRIEVE] [CROSSREF] [CSA]
  • Svane , M. , Hagström , M. and Pettersson , J. B. C. 2004 . Chemical Analysis of Individual Alkali-Containing Aerosol Particles: Design and Performance of a Surface Ionization Particle Beam Mass Spectrometer . Aerosol Sci. Technol. , 38 ( 7 ) : 655 – 663 . [CROSSREF] [CSA]
  • Tafreshi , H. V. , Benedek , G. , Piseri , P. , Vinati , S. , Barborini , E. and Milani , P. 2002a . A Simple Nozzle Configuration for the Production of Low Divergence Supersonic Cluster Beam by Aerodynamic Focusing . Aerosol Sci. Technol. , 36 : 593 – 606 . [CROSSREF] [CSA]
  • Tobias , H. J. , Kooiman , P. M. , Docherty , K. S. and Ziemann , P. J. 2000b . Real-time Chemical Analysis of Organic Aerosol Using a Thermal Desorption Particle Beam Mass Spectrometer . Aerosol Sci. Technol. , 33 : 170 – 190 . [CROSSREF] [CSA]
  • Wang , X. , Kruis , F. E. and McMurry , P. H. 2005a . Aerodynamic Focusing of Nanoparticles: I. Guidelines for Designing Aerodynamic Lenses for Nanoparticles . Aerosol Sci. Technol. , 39 ( 7 ) : 611 – 623 . [CROSSREF] [CSA]
  • Wang , X. , Gidwani , A. , Girshick , S. L. and McMurry , P. H. 2005b . Aerodynamic Focusing of Nanoparticles: II. Numerical Simulation of Particle Motion through Aerodynamic Lenses . Aerosol Sci. Technol. , 39 ( 7 ) : 624 – 636 . [CROSSREF] [CSA]
  • Young , D. F. , Munson , B. R. and Okiishi , T. H. 2000 . A Brief Introduction to Fluid Mechanics. , John Wiley & Sons, Inc .
  • Zelenyuk , A. and Imre , D. 2005 . Single Particle Laser Ablation Time-of-Flight Mass Spectrometer: An Introduction to SPLAT . Aerosol Sci. Technol. , 39 ( 6 ) : 554 – 568 . [CSA]
  • Zhang , X. , Smith , K. A. , Worsnop , D. R. , Jimenez , J. , Jayne , J. T. and Kolb , C. E. 2002b . A Numerical Characterization of Particle Beam Collimation by an Aerodynamic Lens-Nozzle System: Part I. An Individual Lens or Nozzle . Aerosol Sci. Technol. , 36 : 617 – 631 . [CROSSREF] [CSA]
  • Zhang , X. , Smith , K. A. , Worsnop , D. R. , Jimenez , J. L. , Jayne , J. T. , Kolb , C. E. , Morris , J. and Davidovits , P. 2004 . Numerical Characterization of Particle Beam Collimation: Part II. Integrated Aerodynamic-Lens-Nozzle System . Aerosol Sci. Technol. , 38 ( 6 ) : 619 – 638 . [CROSSREF] [CSA]
  • Ziemann , P. J. , Liu , P. , Kittelson , D. B. and McMurry , P. H. 1995 . Particle Beam Mass Spectrometry of Submicron Particles Charged to Saturation in an Electron Beam . J. Aerosol Sci. , 26 ( 5 ) : 745 – 756 . [CROSSREF] [CSA]

Reprints and Corporate Permissions

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

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

Academic Permissions

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

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

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