You are on page 1of 12

J. AerosolSci., Vol. 16, No. 4, pp. M3-354, 1985. 0021-8502/85 $3.

00+000
Printed in Great Britain. PergamonPress Ltd

A NOVEL VIRTUAL IMPACTOR: CALIBRATION AND USE

B. T. CHEla, H . C. YEll and Y. S. CllENG


Inhalation Toxicology Research Institute, Lovelace Biomedical and Environmental Research Institute,
P. O. Box. 5890, Albuquerque, NM 87185, U.S.A.

(Received 4 February 1985)

Abstract--Aerosol separation characteristics of two prototype single-stage virtual impactors are


described. Data indicate that the efficiency curve and the 50 ~o cut-off Stokes number (Stkso) (or cut-
off size) are not sensitive to the nozzle Reynolds number (Re), and can be predicted as long as the
corresponding Re is between 1000 and 8000, the ratio of nozzle to probe distance (S) to the nozzle
diameter (W) is about 1, and the ratio of the hole diameter (D1) to W is between 1.2 and 1.5. The results,
however, show that the cut-off sizes are strongly affected by the ratio of minor flow rate (Q,,) to the
total flow rate (QT). Also, the efficiency data reveal that Stkso of the liquid dioetyl phthalate aerosol is
smaller than that of solid polystyrene Latex aerosol, suggesting internal circulation in the liquid
particles at the acceleration nozzle exists. In addition, we found that, near the cut-off particle size, there
were substantial losses (20 ~ 40 ~o) on the surface of the collection probe. The results also show that
only parameters Dl/Wand Q-/Qr have a slight effect on the shape of the efficiency curve.

INTRODUCTION
There is increasing interest in the use of size-classified aerosol samplers for atmospheric
sampling, because deposition of particles in the respiratory tract and reduction of visibility by
light-scattering strongly depend on particle size. One of the most commonly used size
classification devices is the conventional impactor, originally developed by May (1945), where
particle size fractionation is performed by inertial separation. In the impactor, a small plate is
placed such that the surface is perpendicular to the air flow at the exit of a particle
acceleration nozzle. The performance of a well-designed impactor has a sharp cut-off
efficiency curve following a theoretical prediction (Marple, 1970). However, particle--surface
interaction problems associated with the impactor such as particle bounce, re-entrainment,
break-up, and collection-surface overload have led to non-ideal performance (Rao and
Whitby, 1978; Cheng and Yeh, 1979) and introduced inaccuracy in the data interpretation
(Mercer, 1973).
In search of a technique that uses the concept of inertial impaction, but avoids
particle--surface interaction problems, the virtual or dichotomous impaction principle has
been developed (Hounam and Sherwood, 1965; Conner, 1966; Dzubay and Stevens, 1975;
Forney, 1976; Loo et al., 1976; Willeke and Pavlik, 1978; Masuda et al., 1979; Solomon et al.,
1983). In such a device, the solid impaction plate used in the conventional impactor is replaced
by a region of relatively stagnant air contained in the cavity of a collection probe. The major
air flow which carries low-inertia particles is laterally moved away similarly to the
conventional impactor, while high-inertia particles in the stagnation region are passed into
the collection probe at a relatively low flow rate (minor flow). Both size fractions can be
subsequently ducted to any desired location for further size classification or for any desired
method of analysis, including direct-reading continuous instrumentation.
In this paper extensive experimental evaluation of two prototype round-jet virtual
impactors is presented. In particular, this study documents the effects of nozzle Reynolds
number (Re), minor flow rate, and the geometries of the nozzle and probe on the magnitude
of the particle cut-off diameter, slope of the efficiency curve, and percentage of internal
particle losses. Both solid polystyrene Latex (PSL) and liquid dioctyl phthalate (DOP)
aerosols were used to determine the size separation efficiency of this virtual impactor, but
only DOP aerosols were used to quantify internal particle losses.
343
344 B. "T CHEN et al.

AEROSOL

LAMINATOR- - ~

i
--NOZZLE
SPACER-- '--'SPACER
COLLECTION/
PROBE
MINOR ~ . . ~
FLOW

)
MAJORFLOW
Fig. I. Schematic diagram of the single-stage virtual impactor.

VIRTUAL IMPACTOR
Based on conventional impactor design, two prototype virtual impactors having different
nozzle diameters (W) were designed and constructed. Figure 1 shows the basic design of the
virtual impactor. Aerosol is drawn through an entrance and laminarized by passage through a
ring of small holes (laminator). Flow is subsequently accelerated by the narrowing cross-
section of the duct, thus, increasing the momentum of the aerosol particles. Finally, particles
are separated according to their aerodynamic diameters at the gap between the nozzle and the
collection probe. The major flow which contains most of the fine particles is removed
perpendicular to the acceleration nozzle and collection probe, while the remaining flow
(minor flow) which contains the coarse particles and a small portion of the fine particles
passes across the gap and down through the collection probe. Because of the fundamental air
flow design of the virtual impactor, contamination of fine particles in the minor flow is
proportional to the ratio of minor flow rate to total flow rate. Four pressure taps are used to
monitor the pressures upstream of the nozzle, at the nozzle, and downstream of the nozzle at
both the major and minor flow paths.
Table 1 shows values of design parameters used in the prototype virtual impactors. The
nozzle throat length (T), entrance angle (0), and probe diameter (D2) were identical in both

Table 1. Physical design parameters of two prototype virtual impactors

W (ram) T (ram) 0 (deg) D2 (cm) S (nan) D~ (ram)

Prototype 3.05 3.05 30 2.20 3.15 4.60


I 3.05 3.05 30 2.20 1.63 3.07
3.05 3.05 30 2.20 4.85 4.11

Prototype 2.03 2.03 30 2.20 2.03 2.54


II 2.03 2.03 30 2.20 1.22 2.06
2.03 2.03 30 2.20 3.05 3.07
A novel virtual impactor: calibration and use 345

impactors. Parameters such as nozzle to probe distance (S) and hole diameter (D~)could be
changed to investigate their effects on the size separation efficiency.

METHODS
Generation of solid particles
PSL spheres (Duke Scientific, Palo Alto, California) between 0.6 and 5.0/an in nominal
diameter were generated with a Lovelace nebulizer, dried by a cylindrical diffusion dryer
containing desiccant, and then introduced into a Kr-85 bipolar ion source to bring them to a
state of charge equilibrium (Fig. 2). Before entering the virtual impactor, the particles were
diluted and mixed with filtered compressed air in a mixing chamber.
The number concentrations of PSL spheres in both the major and minor flows were
measured individually by the same Aerodynamic Particle Sizer (AlaS) (TSI, Inc., St. Paul,
Minnesota, U.S.A.) to determine the particle separation effieieneies of both impactors
(Prototype I and II). All the valves downstream from the impaetor were used not only to
adjust the total (Qr), major (QM), and minor (Q,,) flow rates through the impactor, but
also to provide a proper amount ( ~ 5.7 l/rain) (Chen et al., 1985) of air flow through the APS.
Flow rate through a 3-way valve VI to the vacuum was also adjusted to be 5.7 l/rain to match
the total flow through the APS. Figure 2 shows the set-up of valves when the number
concentration of the major flow was measured: V2 and V4 (2-way valves) were closed, Va (2-
way valve) was open, and VI connected minor flow and make-up dilution air to the vacuum.
Normally, both make-up dilution air (Qd) in the minor flow line and bleed-off air (Q~) in the
major flow line were required, because during the ordinary operating conditions (Re ~ 4000),
the minor and major flow rates were less and greater than 5.7 l/rain, respectively. During
operation to determine the concentration of minor flow, V2 was open, Va and V4 were closed,
and V~ was rotated clockwise to connect the major flow to the vacuum. During valve°
switching periods between two measurements, V4 was open to the ambient air to reduce air
flow fluctuations that might be caused by switching valves. The advantage of using one APS
for both major and minor concentration measurements is that the results can be compared
and used to calculate the separation efficiency of the test aerosol without additional
calibration.

, , , , . FILTER
I DIFFUSION ~ BIPOLAR I j MIXING
[ DR;ER I -[ ION SOURCE :I -[CHAMBER~LTOT - L ~

FLOW
L VALVE / 10. N
T
o,LoT,oN 1 v VAOUUM
l I COMPRESSED~
ILOVELACE1 AIR I ~ MAKE-UP J
I I ~ DILUTION I
NEBULIZER Qo
i I V2 F" MINOR , ' ,
A = P, r'.A w I VIRTUAL I

VALVE V1 5.7 I/rain ~CM° U ~ uvv

AMBIENT " ~ 7 ~ V4 ,
I vALvk'-JV 3 1......
/ BLEED-OFF
15.7 I/rain ~ FLOW
T ~ Qb
AERODYNAMIC
PARTICLE
SIZER ~ VACUUM

Fig. 2. Experimental set-up when solid PSL aerosol was used in the study.
346 B T. C u E s et aL

Because the APS has good resolution tbr particles greater than 1 ~m in aerodynamic
diameter (Chen et al., 1985), it was possible to obtain more than one data point per run by
generating PSL aerosols of different nominal sizes concurrently with the same nebulizer

Generation of liquid particles


The purpose of using liquid aerosol in this study was two-fold, first, to compare size
separation efficiency curves for solid and liquid particles, and, second, to determine the
relative amounts of internal losses in the various parts of the virtual impactor.
D O P liquid aerosol in the size range between 1.5 and 16#m were generated with a
vibrating-orifice monodisperse aerosol generator (Model 3050, TSI, St. Paul, Minnesota,
U.S.A.) with an improved liquid feed system (Chen et al., 1984). The feed system provided an
accurate feed rate of a liquid solution containing the desired concentration of D O P dissolved
in isopropyl alcohol with uranine fluorescent dye added as a tracer. The isopropanol/uranine
solution was prepared by mixing 100rag uranine with 1.25 1 isopropanol and filtering to
remove undissolved uranine granules, if any. Solutions with different concentrations of
uranine in the same isopropanol were also used in the early stages of the study and yielded
unsatisfactory results: a solution containing a higher concentration of uranine increased the
viscosity of the D O P solution, resulting in an erratic liquid feed rate as well as the aerosol
generation rate; on the other hand, a more dilute uranine solution took too long ( > 1 hr) to
have high enough uranine deposits for internal wall loss analyses.
The experimental arrangement after the vibrating-orifice generator is shown schematically
in Fig. 3. The aerosol was diluted with clean air and discharged with a Kr-85 source before
entering a manifold. This symmetrical manifold was able to supply aerosol to more than two
samplers. Only one of the virtual impactors (Prototype I) was used to determine the amount
offluorescein deposited on two glass-fiber filters (Gelman Type A/E) in major and minor flow
paths and on various internal parts of the impactor. The APS was used periodically to
monitor the size distribution of D O P aerosols. The degree of monodispersity of D O P was
determined from the spectrum of the APS output.

Vibrating - Orifice
Generator
J
~ Dilution Air

- Overflow
Minor ~
Flow .~ i
Large Particle l%'-'21 I ~ I Aerodynamic1
Co,,ection I Part,o
P' o,.er I
Mai°r FI°w
valve(~ ~ = Vacuum
Small
Vacuum Particle
Collection
Filter
Fig. 3 Experimental set-up when liquid D O P aerosol was used in this s t u d y
A novel virtual impactor: calibration and use 347

Particle separation efficiency


Figure 2 shows the experimental set-up when solid PSL aerosols were used to evaluate the
separation efliciencies of the virtual impactors. For each set of operating conditions (i.e. Re,
W, Qm/Qr, S / W , and DI/W), two number concentration measurements were made for both
the major and minor flows with the APS. The separation efficiency (r/) wa computed from:
~l = C,./(C,. + CMQM/(QM -- Q~)), (1)
where C,,, C~ are mean number concentrations of PSL particles measured by the APS in
minor and major flows, respectively. Equation (1) is valid for most of operating conditions,
except under the cases that R e = 2330 or 1950 with impactor I and R e = 970 with
impactor II when the major flow rate is smaller than 5.7 l/rain. In such a case, bleed-off flow
was replaced by dilution air and tl = C,./(C., + CM).
Liquid D O P droplets (tagged with uranine fluorescent dye) in the major and minor flows
were collected on two filters (Fig. 3) downstream of the impactor (Prototype I). After each
run (30 ~ 50 rain), these filter samples were washed and sonicated separately in 100 ~o
isopropyl alcohol and the fluorescein contents were quantified with a Turner fluorometer.
Efficiency of the separation was calculated from the expression:
tl = m / ( m + M ) , (2)

where m = the amount of fluorescein deposited on minor flow (large particle) filter, and
M = the amount of fluorescein deposited on major flow (small particle) filter.

Wall loss measurements


Internal wall loss studies were performed using liquid D O P aerosols (Fig. 3). After each
run the impactor (Prototype I) was disassembled and the parts (see Fig. 4) were separately
rinsed with 100 ~o isopropyl alcohol. Three aliquots of each washing were quantitated with
the Turner fluorometer. Only the inside surfaces of the impactor were washed. Wall losses in
the impactor were computed from:
W L = M w / ( M w + M +m), (3)
where WL = the fraction of wall losses and Mw = the amount of fluorescein deposited on all
internal surfaces. Wall loss (WL~) at ith part of the impactor was also measured. In this case,
M win the numerator of equation (3) is replaced by Mw~,the amount of fluorescein deposited
on internal surface of ith part of the impactor.

\A/
cl
o11 I
I F ,
I !
Fig. 4. Schematicdrawing of virtual impactor parts that were washed separately to analyze wall
losses.
348 B T CHEN et aL

80

.48
0.91
! 2 . 3 5 2.79

0
I1 .
I I I
129 200 300 384
CHANNEL NUMBER

Fig. 5. Accumulation bin output o f an APS with an aerosol containing, from left to right, 0.91, 1.48.
2.35, 2.79, and 3,30/tin PSL spheres.

RESULTS AND DISCUSSION


Particle separation e~ciency curve
Two aerosols of PSL spheres were generated in this study to determine the separation
efficiency of the impactor, one containing five different nominal sizes (0.91, 1.48, 2.35, 2.79,
and 3.30/an) and the other containing six sizes (0.62, 1.27, 1.74, 2.02, 4.1, and 5,0 #m). Figure 5
shows the output of the multichannel accumulator of the APS for the aerosol containing 0.91,
1.48, 2.35, 2.79, and 3.30/an PSL spheres. Because no overlapping among the peaks
representing different sizes was observed in the output, these two aerosols were used
alternately to obtain either five or six data points per run. Some of the liquid DOP ~ t i c l e s
were also produced to investigate potential differences in the particle separation efficiency
curves for solid and liquid aerosols.
As discussed by Marple and Chien (1980), the flow field, particle trajectory, and the
corresponding ©ffmiencycurves of virtual impactors depend on the flow parameters, nozzle
Reynolds number (Ri) and minor flow rate ratio (Q-/Qr), and the physical design
parameters, dimensionless nozzle to probe distance (S/W) and dimensionless probe diameter
(D~/W) (both nozzle throat length, T, and entrance angle, 0, are fixed in the impactors). To
maintain a parametric study throughout the experiments, a set of base values for the
parameters was chosen and listed in the first row of Table 2 for each impactor. The other
values of the parameters in Table 2 were then varied one at a time, while the remaining
parameters remained fixed. The following are results and discussions of the effects of various
parameters on the size separation efficiency curves.

(i) E~ect of Re
To determine the influence of Re, experiments were run with the Re listed in Table 2, while
the other parameters were held constant. Separation efficiency curves of virtual impactors I
and II as a function of aerodynamic particle size (D~E)of PSL particles are shown in Figs 6(A)
and (B), respectively. Similarly to the characteristics in a conventional cascade impactor, the

Table 2. Operating conditions and design parameters used in the study

Re Q,/QT (%) S/W Dt/W

Prototype t 3900 + 50 10.8 + 0.9 1.03 1.51


I 1950, 2330 6.0 0.53 1.01
(W= 3.05 mm) 8180, (7400)* 14.0, (12.5, 18.1)* 1.59 1.35

Prototype~ 4200 + 50 10.9 + 0.9 1.00 1.25


II 970 3.0 0.60 t.01
( W = 2.03 mm) 7920 22.0 1.50 1.51

* The values in parentheses represent the conditions when DOP-particfes were


generated for internal loss analyses.
~ The values in the first row of each impactor are the base values.
A novel virtual impactor: calibration and use 349

EFFECT OF Re
VIRTUAL IMPACTOR I

7P- 100
uJ
/ [~Re=3900±50
ee
Lkt
A
~Re:2330
~.. 8O
Re=1950

~ 50
7
al

U-
ii 40
Z
0
7- 20
<
er
<

00.5 1.0I I
2.0
I I . I
4.0 5.0 10,0
1
A E R O D Y N A M I C DIAMETER (DAE), p.m

EFFECT OF Re
VIRTUAL IMPACTOR II
,00 8o
B
Sly oo

i 60 =~Fi
970e=
40
ktJ

20

I I l I I I
t.u
oo 0 0'.5 1.0 2.0 4.0 5.0 10.0
AERODYNAMIC DIAMETER (DAE), film

Fig. 6. Separation efficiency as a function of aerodynamic diameter of PSL spheres for (A) impactor
I(Q=/Qr = 10.8 + 0 . 9 9 o, S/W= 1.03, and DI/W= 1.51) and (B) impactor II (Q,/Qr= 10.9 :i: 0.9 9o,
S/W= 1.0, and Dt/W= 1.25) operated under different nozzle Reynolds numbers (Re).

data indicate that with a fixed nozzle diameter (W), a virtual impactor operated at higher Re
has a smaller 50 ~o cut-offaerodynamic diameter (Dso), and with a fixed Re, the impactor with
a larger W(Impactor I) has a larger Dso. We expected the efficiency curves for different Re to
be asymptotic to about 11 9o for smaller particles, because contamination of the fine particles
in the minor flow is equal to Qm/Qr, which is about 11 ~ for the operating conditions of both
impactors. The variation of efficiencies (Figs 6A and B) between O and • in both (A) and (B),
and between [] and • in (A) are mainly due to small variation in Q,~IQr value.
All the data shown in Figs 6(A) and (B) can be consolidated into a single curve (Fig. 7) ifDAE
is replaced by a dimensionless particle diameter, the square root of Stokes number (Stk),
which is defined as
Stk = D~ECVo
9~W "
where C is the particle slip correction factor, V0 the mean fluid velocity at the nozzle throat,
and/a the fluid viscosity. And the (Stk)~82 value corresponding to 50 ~o separation efficiency in
both impactors is about 0.67. The results shown here suggest that we could predict both the
separation efficiency curve and the Dso of any geometrically and dynamically similar virtual
350 B. T CHEN et at

100
z
UJ Re=7920 _ _ _ ~ = 3900*.50
o Re=4200 *-.50--- ~:~
rr
Lu 80
13... ~--Re=1950
Re:4200±50 -~ .~
~- 6oi Re=970~ Re=3900±50
o
7, ....... Re=2330
_c2 40
IJ.
7 Re 8180
o 2o
rr
<
O- I I I I i I
00.1 0.2 0.4 0.5 1.0 2.0
( S t k ) 1/2

Fig. 7. Separation efficiency as a function o f the square root of Stokes n u m b e r (Stk). Data from
Figs 6(A) and (B) are represented by a single hand-fitted curve.

impactor of fixed Q=/Qr if Re is between 1000 and 8000, S/W is ~ 1, and D1/W is between
1.25 and 1.5.

(ii) Effect of Q,./QT


Particle separation etficiency curves corresponding to different Q,/Qr values are shown in
Fig. 8. Results obtained with solid PSL particles indicate that the 50 % cut-off size (or
(S tk )~/oz) decreases as Q,/ Qr increases. This is expected, because more str~_ mlir~es containing
more small particles pass through the collection probe for larger values of Q,,/Qr (Marple
and Chien, 1980). It is also int~esting to note that the changes in Q,/QrValU¢ do have some
effect (especially in Prototype I) on the slop~ of the etl~ienes, curves, because the curves are
asymptotic to the different valtms of Q=/Qrat low values of (Stk) 1/2. Data points for D O P
particles (Fig. 8A) also have smaller cut-off size and smaller slope as Qm/Qr increases. The
50% cut-offs (Stk) ~/2 are about 0.5 and 0.45, for D O P at Q,~/Qr = 12.5 and 18.1 ° o,
respectively, and about 0.75, 0.67, and 0.6 for PSL particles at QM/Qr = 6.0, 9.9, and 14.0%,
respectively. The difference of ($tk)~/o2 for solid and liquid aerosols sugi~ts that the D O P
droplets, though they are spherical, undergo an internal circulation and a surface motion
at the acceleration nozzle, resulting in a smaller drag co¢tficient and, subsoqutmtly, a larger
apparent size than those of solid particles (Cliff et al., 1978). As a result, the actual
aerodynamic sizes and Stokes numbers should be larger than those used in the figure and the
D O P data should be shifted to the right.
Figure 9 shows the rdatiouship between Q=/Qrvalues and the corresponding (Stk)~/o2 for
PSL aerosols. The curve fit was determined by least-squares polynomial regression using all
the data:
(Stk)~/oz = 1.2766 - 0.8081 (log~o (Q,,/Qr) )
+ o.1922 (logl o (Q,/ Qr ) )2.
This curve can be used to predict 50 % cut-off (Stk) ~/2 values for Q,,/ Qr values betw~n 0.03
and 0.25. It is important to note that varying the Qm/Qrvalue in the range greater than 0.3
may cause a substantial change in the value of (Stk)~/o2 because (Stk)~ 2 approaches 0 as
Q,,/Qr closes to 0.5.

(iii) Effect of S/W


The influence of S/W was determined by using values of S / W = 0.53, 1.03 (base), and 1.59
for impactor I and values of S / W = 0.60, 1.00 (base), and 1.50 for impactor II (Fig. 10). Only
A novel virtual impactor: calibration and use 351

EFFECT OF Qm/Q T
VIRTUAL IMPACTOR I

t- 100
Z A
LtJ
O
LU 8O
g--
v
>- 60 Qm/QT •
o
z
tu ~.~-<~>--~-"/4 /
,Sld
t~
I.L 40 ,,.,,<oo,,--.
ILl
z
o 20
C-- ~'- 6.0%(PSL)

m
n I I I I i
IJJ 0 I 1 I I
0.1 0.3 0.5 07 1.0
(Stk)~12

EFFECT OF Qm/QT
VIRTUAL IMPACTOR II
100
B
~_ 8o

°"°----' IJ l

°77 i
60

E 4o
12J ~.°,,---2~ ? _/
!2o
, ~nim"~4n's' i 1,11 ° 1
~ 00.1 0!2 21.0
(Stk) 112

Fig. 8. Separation efficiency curves showing the effect of Q - / Q r in (A) impactor I (Re = 3900 + 50,
S / W = 1.03, and DI/W = 1.51) and (B) impactor II (Re = 4200 + 50, S~ W = 1.00, and DI /W = 1.25).

1 .0~

II
C•Prototype
"~= 1.2766-0.808 1 ( I O g l o X)
0.8 ~ + 0 . 1 9 2 2 ( I O g l o X) 2
tZ3 Prototype I- - - - - ' @ ~

0.6~-

0.5 i l I I I
2 3 5 10 20 30
Qm/Q T ( P e r c e n t )

Fig. 9. Et']['CCIo[" (2,,1/QT Oil 50o0 CUt*ON (Slk) 1'2 Of the virtual impactors.
352 B 'I~ CHEN et aL

EFFECT OF S/W EFFEC 1 ,;DF S, v,+


VIRTUAL IMPACTOR I VIRTUAL i",APACTOF,: I

100
Z

A ~~ IO0 If B J
W
~ 80 a_ go
"V /
~
~ 60 , >. /
I
Z
w D ~ 1.03 (-) S/W ]
i
t
~ 40- j .... 0.53 -- 40
U-
h
060
- --4/
W W 1 50 ........ ~ 7
l
Z -- Z oo ..... -:7
0 _o 2o
~ 2O
<

1 I I I I I i I.iJ I l
w 0
m 0.1 0.2 0.4 0.5 1.0 (,,) O0 1 0.2 ' o.'4o;s' ' ' io
( S t k ) ~j2 (Stk)~'2
Fig. 10. Selmration efficiency curves showing the effect of S I W in (A) impactor I iRe = 3900 + 50,
(2=/0. = l O . 8 ± 0 . 9 % , a n d D i / W = 1.51)and (B)imlm~or II (Re = 4200±50,Q.,/QT= 10.9±0.9",.
and Dr~W= 1.01).

PSL particles were used. The results shown in Fig. 10(B) for impactor II are very similar for
three different $/W values and are represented by a single curve (Dr / W value used here is 1,01
instead of the base value 1.25). Figure 10(A), however, shows a slisht difference between the
curve for S/W= 0.53 and that for $/W= 1.03 and 1.59. Despite the indicated difference, all
the data here suggest that the effect of S/W on virtual impactor performance is small
compared with S/W effect for conventional impactors (Marple and Liu, 1974).

(iv) Effect of D1/W


Figures 11 (A) and (B) show the effect of D1 / W on the separation efficiency curves for PSL
particles. Both figuros show that the efficiency curves for Dl / W = 1.01 have steeper slopes and
smaller cut-off siz~ than those for other larger Dt/W values, indicating better cut-off
characteristics. Also, the particle selmmtion curves are almost identical for Dt/W values
greater than 1.35 and 1.25 in impactor I and II, respectively. It is important to note that

EFFECT OF Dj/W EFFECT OF D/W


VIRTUAL IMPACTOR I
100 f._ 1 0 0
Z
UJ A D/W
z
w B
(3 0
1.35-- ----/ I~
w 80 w 80

U
n

g-
v
>- 60 >. 60
(3
Z 1.51- J Z
O
W
ua

,7 40 _o 4O
LL
W
Z
o_ 2o _o 20
I-- I,.-
<
he
< <
n
l I I I I I Ii,,I I I l I 1 I I I
w 0 I Ill 0 ~ 0
o9 0.1 0.2 0;4 0 5 . ~0 0.1 0.2 0.4 0.5 1.0
( S t k ) lt2 ( S t k ) w2

Fig. 11. Separation efficiency curves showing the effect of Dl / W in (A) impactor I ( Re = 3900 + 50,
Q,-/QT = 10.8 + 0.9 ?'o, and S / W = 1.03) and (B) impactor II (Re = 4200 4- 50, Q,,/QT = 0 9 + 0 9 '~,
and S / W = 1.00).
A novel virtual impactor: calibration and use 353

selection of larger Dt/Wvalues ( > 1.01) as base values is based on a trade-off, i.e. wall losses
are lower for larger values of Dt / W, though the cut-off characteristics might not be as good.

Wall losses
Liquid D O P particles tagged with uranine were used for internal wall loss analyses in
impactor I. Figure 12 shows total particle losses inside impactor I at three different operating
conditions ( S / W = 1.03 and D I / W = 1.51 were fixed), and Figure 13 shows wall losses
measured for individual parts (Fig. 4) of the impactor. For all conditions, the highest losses,
mainly on the upper surface of collection probe, are shown at the sizes approximately
corresponding to the 50 % cut-off (Stk) ~/2. For particles larger than 10/am in aerodynamic
diameter (i.e. (Stk) ~/2 > 2.0 and 3.0 in the cases of Re = 3850 and 7400, respectively), the wall
losses were high, primarily due to impaetion on the inner surfaces of the converging region of
the acceleration nozzle. It is interesting to note that around the cut-off sizes, losses at
Re = 3850 increase as Q~,/Qrdecreases, and losses decrease with increasing Re. These results
agree with those presented by McFarland et al. (1978). Figure 13 shows typical losses on
internal parts (A) nozzle, (B) collection probe, (C) spacer, and (D) transport of major flow (see
Fig. 4) of the impactor I operated at Re = 3850, and Q,,/Qr = 12.5 ~o. The amount of losses
due to parts (E) and (F) are negligible.
Wall losses on the upper surface of the collection probe of this prototype virtual impactor
were large (20 ,-, 40 ~ ) for particles near the cut-off size (Fig. 12), indicating that the hole

VIRTUAL IMPACTOR I
40

Re Qr./QT
3o ~ 3~o
3850 18.1%
7400 12.5%
Q_
~ 20

-~ ~o

0 . I
0.3 0.5 1.0 2.0 4.0 5.0
(Stk)~ ~2

Fig. 12. Total wall loss versus (Stk) 1~2 for virtual impactor 1 (S/W= 1.03, a n d DI/W= 1.51).

VIRTUAL IMPACTOR I
40 Re Qm/QT
385--0 12?5%
C ollectio, n P r o b e
Z 30

cO" 2 0
0
,m,,o, ,,ow,
.J
~ 10

I Nozzle

0.3 0.5 1.0 2.0 4.0

( S t k ) ~j2

Fig. 13. Wall losses measured for individual parts of virtual impactor I, labelled as in Fig. 4
(S/W= 1.03, a n d DI/W= 1.51).
,354 8 ][ CHENet al

diameter o f the p r o b e m a y be t o o small and, as a result, some particles near the cut-off size
impact and are deposited on the surface (like the impaction plate o f a conventional impactori.

CONCLUSIONS

T h e present study has experimentally confirmed that these p r o t o t y p e one-stage virtual


impactors can be used as size-selective samplers with different cut-off sizes. The particle
separation efficiency curve is c o m p a r a b l e in sharpness to that o f a conventional impactor, but
is not identical for solid a n d liquid aerosols, because o f the existence o f internal circulation in
liquid droplets at the acceleration nozzle. Results showed that the 50 9o cut-off Stokes
n u m b e r (Stk) is substantially affected by the ratio o f m i n o r flow rate to the total flow rate,
Q m / Q r , but n o t by the nozzle Reynolds number, Re. Separation efficiency data for R e from
1000 to 8000 fit a c o m m o n curve when plotted versus the (Stk) t/2. Therefore, we conclude that
the cut-off size can be predicted if the i m p a c t o r is operated with R e in the range between 1000
a n d 8000. This assumes that the data for the dimensionless parameter (Stk) t/2 can serve as a
basis for design o f d i c h o t o m o u s samplers for other cut-off sizes. In addition, with a fixed R e
the cut-off size is smaller with smaller nozzle diameter
The results indicate that the separation efficiency curves have essentially the same shape
and are not substantially influenced by any o f the parameters tested. O n l y two parameters,
Q m / Q r and D t / I g , a p p e a r to have a slight influence on the shape. F u t u r e work will
concentrate on the reduction o f internal particle losses. Also, a p r o t o t y p e two-stage virtual
i m p a c t o r will be evaluated to p r o d u c e aerosols with n a r r o w a e r o d y n a m i c size ranges.

Acknowledgements--The authors are indebted to many of our colleagues for technical review and suggestions;
Drs B. B. Bocckcr, C. H. Hobbs, and R. O. McClellan for encouragement and support; Mr F. J. Greybar for
constructing the impactors; Mr E. E. Goff for illustration; and Word Processing porsonnel for typing. This research
was performed under U.S. DefraYment of Energy Contract Number DE-AC04-76EV01013

REFERENCES
Chen, B. T., Cheng, Y. S. and Ye2a,H. C. (1984) J. Aerosol Sci. 15, 457.
Chen, B. T., Cheng~ Y. S. and Yeh, H. C. (1985) Aerosol Sci. Technol. 4, 89.
Cheng, Y. S. and Yeh, H. C. (1979) Environ. Sci. Technol. 13, 1392.
Clift, R., Grace, J. R. and Wel~r, M. E. (1978)Bubbles, Drops and Particles, pp. 97-141. Academic Press, New York.
Conner, W. D. (1966) Air Pollut. Control Assoc. J. 16, 35.
Dzubay, T. G. and Stevens, R. D. (1975) Environ. Sci. Technol. 9, 663.
Forney, L. J. (1976) Rev. Sci. lnstrum. 47, 1264.
Hounam, R. F. and Sherwood, R. J. (1965)Am. Ind. HYO. Assoc. J. 26, 122.
Loo, B. W., Jaklevic, J. M. and Goulding, F. S. (1976) Fine Particles (Edited by Liu, B. Y. H.), pp. 311-351. Academic
Press, New York.
Marple, V. A. (1970) Ph.D. Thesis, University of Minnesota.
Marple, V. A. and LiLt, B. Y. H. (1974) Environ. Sci. TechnoL g, 648.
Marple, V. A. and Chien, C. M. (1980) Environ. Sci. Technol. 14, 976.
Masuda, H., Hochrainer, D. and St6ber, W. (1979) J. Aerosol Sci. 10, 275.
May, K. R. (1945) .l. Sci. lnstrum. 22, 187.
McFarland, A. R., Ortiz, C. A. and Bertch, R. W. Jr. (1978) Environ. Sci. Technol. 12, 679.
Mercer, T. T. (1973) Aerosol Technology in Hazard Evaluation, pp. 222-240. Academic Press, New York.
Rao, A. K. and Whitby, K. T. (1978) ./. Aerosol Sci. 9, 87.
Solomon, P. A., Moyers, J. L. and Fletcher, R. A. (1983) Aerosol Sci. Technol. 2, 455.
Willeke, K. and Pavlik, R. E. (1978) Environ. Sci. Technol. 12, 563.

You might also like