You are on page 1of 14

OPTICAL TRANSMISSION METHOD FOR DETERMINING SOOT

CONCENTRATION IN PLUMES
by Oscar Biblarz

ABSTRACT
In this work, a new and more direct method to resolve optical-transmission
measurements to determine soot concentrations (Cm) is presented. When combined with
an appropriate exit flow velocity and area, the procedure yields the soot aerosol flow rate
(or mass loading) issuing from a gas turbine or rocket. In the determination of aerosol
density, no intermediate Sauter-mean-diameters are required, only a material reference
parameter that distinguishes soot from other particular suspensions. Also avoided is the
explicit use of the complex index of refraction that for soot is associated with high
uncertainties. In order to analyze any transmission measurements, there is a need to
arrive at averages over particle size but this averaging is shown to be only weakly
dependent on the form of the probability density function (PDF) used. Our method can
be both accurate and practical since existing multiple-wavelength transmission
measurement methods may be applied, and the technique is non-intrusive, in-situ, and
appears to be suitable for “hardening” in test-cell environments. The new method
presented here describes available data on Cm within 15 % over a very wide range of
experimental conditions.
INTRODUCTION
Optical-measurement instruments can provide for useful, near real-time
evaluation of particle concentrations in plumes. They are nonintrusive and feature high
sensitivity with rapid response. The measurable quantities comprise both light absorption
and scattering. Most current techniques measure light extinction by particles in the path
of a highly collimated, monochromatic light beam from a laser. Present instruments,
however, rely on a precarious blend of theory and experiments where key uncertainties
are difficult to surmount. Furthermore, it becomes apparent that many of these
instruments comprise relatively impractical hardware for use outside of the laboratory.
This work describes how optical transmission measurements (i.e., extinction) may
be used more directly to obtain soot aerosol densities. When properly combined with the
exit flow velocity and area, they yield the soot mass flow rate (or mass loading) issuing
from gas turbines or rocket engines. In the determination of aerosol densities, our
technique requires no additional measurements or intermediate calculations of particle
diameter, except for the use of a material reference parameter that distinguishes soot from
other particulate suspensions. The use of probability theory is minimal and only
relatively simple algorithms are needed for the calculations. In this paper we report on
the accuracy of the results as well as on the relative affordability aspects for the
applications envisioned.

1
BACKGROUND
Present nonintrusive methods (1-5) for determining aerosol densities depend on
the definition of a Sauter-mean particle diameter that is derived from a blend of Mie
theory with probability theory. An equivalent spherical diameter together with a particle
mass density yield an average particle mass which when combined with estimates of the
number of such particles per unit volume results in the desired aerosol density mass
loading. Mie theory, strictly speaking, applies only to single spheres and requires precise
knowledge of the complex index of refraction of such spherical particles. Probability
density functions are purely empirical, usually obtained from sampling techniques that
are known to distort the aerosol composition being measured. Furthermore, soot
suspensions in plumes exist in a turbulent environment with prevailing non-equilibrium
and polydispersion. Soot particles, moreover, have a strong tendency to agglomerate (6,
7, 8) producing non-spherical shapes of fractal dimensions that reflect very complicated,
non-spherical shapes (9). Thus, the application of Mie theory together with an a priori
probability density function (PDF) can only be expected to yield rough estimates of the
mass loading.
Independent scattering measurements are sometimes used to complement
extinction measurements. As a rule, scattered signals are typically much weaker than
those measured by transmission (2-4, 10, 11). Scattering is also quite sensitive to
agglomeration, particle shape and polarization of the light beam, all of which complicate
scattering measurements considerably (but can be ignored in transmission measurements
(10, 11)). In sooting environments, much of the scattered energy is in the forward
direction falling into the transmitted beam (thus contributing to extinction). Directions
other than forward are necessary for the measurement of purely scattered radiation
requiring a more complex apparatus. In addition, large-angle scattering instrumentation
is far more difficult to operate, particularly in the real environment of engine test cells.
TECHNIQUE FOR DIRECT MEASUREMENTS
Transmission measurements have been fast evolving into standardized techniques
in the field of optical diagnostics of combustion particulates (2-4). In addition to the
advantages enumerated above, such instrumentation can be relatively portable which is a
necessity for in-situ measurements. Fast microprocessors can read and interpret the data
so that near real-time results can be obtained with simple algorithms. Single and multi-
wavelength lasers or discharge lamps provide for intense, controllable sources of
electromagnetic radiation. Intensities before and after passing through the test volume
(from a rocket or a gas turbine exhaust) are usually measured with appropriate off-the-
shelf detectors. A transmission ratio may thus be established where care is taken to
exclude spurious scattering so that purely beam extinction measurements are made (1).
Two related parts need to be developed for the realization of a useful soot-
determining field instrument. The first one is a simpler but accurate procedure for
arriving at the aerosol density based solely on transmission measurements, and the
second one is an improved configuration for obtaining such transmission measurements.
Both of these aim at facilitating any task leading to a practical instrument.

2
Rayleigh Regime in Mie Theory Soot particulates range in size (d) from as low as 5
nm to perhaps as large as 300 nm, but the bulk of the mass is contained in the smaller
sizes so that the dimensionless parameter x = πd/ λ may be treated as less than one in the
visible spectrum. Thus, most of the particles are small compared to the wavelength (λ) of
the probing light (in the Rayleigh regime). The transmission T(λ) is defined as the ratio
of the intensity transmitted to that originating from the source. From Mie theory (12)
T(λ) = exp (- QeANLe) (1)
QeA = Qe (¼πd2) ≡ qe (2)
Here Le is the extinction pathlength. The parameter qe represents the extinction cross
section, Qe the efficiency factor, N the number density of soot particles, and A the
physical cross section. The extinction coefficient, αe, according to Lambert-Beer’s law is
found from
αe = qeN (3)
hence, T = exp (- αeLe) (4)
An new “extinction parameter” (13-16) for a single component submicron
suspension of diameter d may be defined as

qN Q (π d2 / 4)N 3Qe
µe ≡ e
= e = (5)
Cm ρ p(π d / 6)N 2ρ pd
3

where Cm is the desired soot concentration and ρp the particle density. Since Qe depends
on the index of refraction, m = n – ik, then μe depends on λ, d, m and ρp. According to van
de Hulst (12), in the Rayleigh regime the efficiency factor becomes

m2 − 1
Qe ≈ − I m(4 ) (6)
m2 + 2
After some manipulation, the expression for μe turns out to be,

6π 6nk
µe = [ 2 2 ] (7)
ρ pλ (n − k + 2)2 + 4n2k2

which is independent of particle diameter, particle shape and particle size distribution. In
other words, from Eq. 7, μe is only a function of wavelength and of a material property.
A polynomial dependence in λ should be sufficiently accurate to represent μe since in
actuality it is only weakly dependent on wavelength as shown in Fig. 1 (which is a
composite of data from Refs. 13-16 representing measurements from the seeding of
carbon powders, of size 0.012 and 0.025 microns established prior to agglomeration, into
a controlled flow). Details of the polynomial will be given in the next section. It can also
be shown (12) that in this case extinction derives mostly from absorption since pure
scattering represents less than 1 % on the signal intensity.

3
Equation 7 may be used to perform an uncertainty analysis (17) on μe. If we
consider the individual uncertainties in ρp, n, and k, we find that any calculated values of
μe can vary by a factor of two based on known values. Moreover, the largest contributor
is by far the complex index of refraction, k, which is unfortunately poorly known from
the literature. Figure 2 shows curves of Q/d versus d where a change in k results in a
nearly proportional change in Q/d for fixed n, d and λ. For this reason, the establishment
of the relevant extinction parameter under a particular experimental condition is very
desirable and this is discussed in the next section. Also, it is imperative to choose a
fiducial value that minimizes uncertainties in μe.
Lorenz-Mie Theory
In actuality the Rayleigh regime is not completely accurate for soot suspensions
when visible light is utilized as can be seen from Fig. 2. In this figure, the full Mie
calculations are shown for Q/d as a function of d. According to Eq. 5, μe is directly
proportional to Q/d since all other terms are constant. As can be seen, at first the curves
slope upwards with increasing diameter then they peak and then drop displaying a
complicated dependence on the diameter. Moreover, the calculation scheme outlined
below depends only on measurements that are taken at several wavelengths but which by
their very nature average over particle diameters. In actuality, the extinction coefficient
reflects also the Sauter-mean-diameters (1, 10, 11), d32, but in complicated ways. It is
necessary, therefore, to suitably average the equation for Cm so as to properly represent
the nature of the measurements. This necessitates the use of a particle size distribution
function through the probability density function (PDF).
The most commonly used PDF relating to aerosols is the log-normal distribution.
Dobbins (10) reports on the form of d32 as well as on other moments of this distribution.
Since for spheres there is a cubic dependence of Cm on the diameter (see Eq. 5), the
appropriate average is d30 and the relation between these two moments of the size
distribution is readily found as
d30/d32 = exp(3/2 σ2]/ exp(5/2 σ2] (8)
Here, σ is the ‘natural log’ of the geometric standard deviation σg. For a value
representing soot of σ = 0.405 ((10), or σg = 1.5 (1)), the ratio given by Eq. 8 turns out to
be 0.611. It is interesting to note that a uniform distribution spanning just the expected
diameter range of the soot actually yields a ratio of 0.593 so that the result above is fairly
independent of the PDF’s exact shape. Thus, Eq. 8 reflects a sensitivity to the range of
diameters in the PDF rather than any skewness or other characteristic.
Now, the relation between averaged properties (with <>) based on d32 can be
written in terms of β, a numerical constant as yet undetermined,
<Cm> = β<αe>/<μe> (9)
It will be assumed hereafter that the contribution to β from Cm is 0.6. The value of <αe>
must be derived from measurements that directly reflect d32, and contributions to β from

4
<μe> may be found from Mie theory which is rather complicated, but fortunately this
contribution is expected to be minimal.
Method of calculation
Spectral transmission measurements are routinely made at several wavelengths
because conventional theories are based on the ratios of these measurements (1-4).
Three-wavelength measurements are most often used (1), but more than three have also
been employed (2-4). The acquisition of these data is nearly simultaneous and
experimental uncertainties can be identified, although some assumptions leading to the
multi-line calculations still remain debatable.
The method of approach we use is based on the previously given analysis and on
the recognition that the extinction parameter <μe(λ)> as it relates to soot aerosols remains
sufficiently constant with variations of temperature and pressure in ranges of practical
interest (13-16). Implicit here is the fact that real soot-aerosol particle configurations
need not be explicitly represented with respect to shape, size, or any other intermediate
physical property. In fact, the empirical value of the absorption coefficient reported in
Refs. 15-16 is <μe> ≈ 5,000 m2/kg between 300 and 700 nm and for temperatures up to
2200 K, and with a variety of pressures. We may take this equality to be most
representative at 450 nm where the uncertainty in k appears to be a minimum (2-4, 6, 10,
11) and assume a polynomial of the form
<μe(λ)> = a + bλ + cλ2 (10)
There is no special advantage to finding n and k explicitly because they will reflect much
variation under the many different experimental conditions commonly encountered. The
polynomial in Eq. 10 may be evaluated by taking transmission measurements at three
different wavelengths, say <T(λ1)>, <T(λ2)>, <T(λ3)>. The ratios (b/a) and (c/a) are
obtained from two simultaneous equations
(λ1 – λ2 R12)(b/a) + (λ12 – λ22 R12)(c/a) = (R12 - 1) (11)
(λ1 – λ3 R13)(b/a) + (λ12 – λ32 R13)(c/a) = (R13 - 1) (12)
where R12 = ℓn(1/<T(λ1)>)/ ℓn(1/<T(λ2)>) and R13 = ℓn(1/<T(λ1)>)/ ℓn(1/<T(λ3)>).
Standard matrix techniques can be used to solve the corresponding (b/a) and (c/a) ratios
for soot. Note that when taking ratios Cm cancels out because it is λ-independent. The
remaining constant “a” is obtained from the empirical information for soot,
a = 5000/[1 + (450)(b/a) + (450)2(c/a)] (13)
This then defines the constants in the equation for <μe(λ)> according to the particular
experimental situation at hand.
Now, to arrive at the aerosol concentration or density (<Cm>) we manipulate the
following relations,
μe = (qe/mp)(N/N) = αe/Cm (4)

5
<T> = exp(-<αe >Le)
Equivalently,
<αe > = (1/Le) ℓn (1/<T>)

So that with the introduction of Eq. 9 we get

β 1
< Cm >= l n( ) (14)
< µe(λ ) > Le < T (λ ) >

We have introduced earlier Le, the extinction pathlength, and <αe(λ)>, the
extinction coefficient as given in Lambert’s law. The parameter qe/mp represents the
extinction cross section divided by the particle mass, and N represents the number density
of particles. Recall that extinction is more than pure absorption because it includes the
forward scattered radiation (12).
Figure 3 depicts a compendium of transmission data from several authors. These
data are reduced both with the traditional Mie-theory technique that incorporates
intermediate particle size and probability theory (Method I), and one using the procedure
detailed in this paper (Method II). The soot data presented were obtained in four
different laboratories (1, 2-4, 18-21) and represent a very wide variety of experimental
conditions. It is interesting to note that within a definable error margin and over nearly
four-decades of magnitude, all results closely correspond. Thus we may surmise that
with Method II both the trend and the magnitude are accurately represented. Except for
Ref. 16, the original experimental sources do not report on independent means of
verifying Cm. Thus, it is not possible to further identify the veracity of the reported data.
Contributions to β of a suitably averaged μe may further modify the value of this constant.
Our independent examination of 66 data points shows that the ratio of β/<μe(λ)> is
indeed reasonably constant with an average value of 9.7x10-5 kg/m2. This value may be
used to refine the entries shown in Fig.3 shown as Method II since results appear to differ
by small constant offset.
There are several good reasons for the retention of multiple-wavelength
measurements. First, they allow for a verification of the trueness of the data; a single
measurement of transmission at 450 nm in not adequate because of the need to establish
if the experimental situation really represents purely soot (i.e., a single component
aerosol); for example, our technique has been applied to reduce transmission data where
large, non-soot particles were present engendering a bimodal distribution (4) and the
resulting densities do not show any consistency (e.g., 2.14x10-4 kg/m2 at 250 nm,
6.34x10-7 kg/m2 at 350 nm, and 6.64x10-3 kg/m2 at 450 nm). The three calculated
densities need to be much closer to each other, perhaps within 10 %, in order to reflect
the trueness of the data. An average of the resulting aerosol densities can always be
performed to further minimize measurement errors. When there is no bimodality, our
calculations indicate that Method II is, in all cases, equivalent to the more complicated
calculations in the literature.

6
A second advantage of the multi-wavelength approach is that it allows for the
exploration of more than one reference value. The sensitivity to the choice of <μe(λ)> =
5000 at 450 nm has been examined using data from Ref.1 and we found that the given
calculations represent the least variation in density (actual density values are: datum 450
nm, 3.20/3.18/3.32 or a 4% variation, datum 630 nm, 1.85/1.84/1.95 or a 6 % variation,
datum at 1000 nm 0.86/0.82/0/86 or a 5 % variation).
Total Mass Flow Rate
Aerosol flow rates can now be related to the total engine mass flow rates. Engine
flow rate found from the known fuel and air flow rates; this also establishes gas
composition. Instead of measuring velocity and area, all that needs to be established is
the average gas temperature (<τ>). Then, the perfect gas law may be used to determine
the density (ρg). The following equations apply:
ρg = p/R <τ> (15)

VA = m
&t / ρg (16)

&= m
m &t (<Cm>/ ρg) (17)

where m &t and m&represent the total and the aerosol mass flow rates respectively, V and A
are the exhaust velocity and the area respectively, p the exhaust pressure, and R
represents the exhaust-gases gas constant. This procedure works well when the particles
being measured are distributed reasonably homogeneously as in turbulent flows.
INSTRUMENTATION
An improvement in the standard way of measuring transmission (2-4,18-21) may
be realized with the introduction of a device that can accurately redirect be beam back
through the test volume toward the light source. This simplifies alignment, eliminates the
need for fiber-optic cabling, and doubles the extinction length. Increasing the extinction
length is quite desirable when the light source is sufficiently intense and the plume itself
is a weak attenuator. Figure 4 depicts a configuration where only one of several lasers
(of different wavelengths) is shown. This set-up is an adaptation of the hardware in Refs.
2-4. The light chopper (optical or mechanical) is necessary to establish <T> (which is
found from the ratio of measured intensities, I/Io, I being the transmitted intensity and Io
the reference intensity or intensity without particles). Since local environmental
conditions (vibration, heating, etc.) cause fluctuations in the light intensity, I and Io must
be nearly simultaneously recorded and with the same diode. The experimental
arrangement depicted allows measurements of the sum of the two intensities where one is
chopped at high speeds. Beam steering arising from changes in index of refraction across
the hot plume can be the source of some error and should be scrutinized in each and
every experimental arrangement.
One candidate device for accurately redirecting the beam is the so-called
retroreflector (22, 23), see Fig. 4, also know as a corner-cube reflector or trihedral prism.
This unit consists of s system of 3 fixed mirrors so arranged as to reflect back light
7
coming from directions within a broad acceptance angle. High quality retroreflectors are
commercially available with accuracies from 5 minutes to one second of parallelism, or
to 1/4–wavelength of light in flatness. Mirror surfaces can be made to efficiently reflect
wavelengths from the infrared to the ultraviolet. Preferably, the returned beam should be
displaced from the source so as to simplify the optics. Retroreflectors are relatively
inexpensive, rugged and commonly available in appropriate configurations.
SUMMARY, CONCLUSIONS AND RECOMMENDATIONS
One singular advantage of our proposed scheme is that aerosol densities are
obtained more directly and rapidly from transmission measurements, simplifying the
calculational requirements. No intermediate particle diameters are necessary; a
probability distribution function is only needed to establish the constant that goes into the
calculation of Cm (Eq. 9). As such, the scheme is simpler than conventional approaches
and it avoids any strong reliance on parameters that have large uncertainties such as the
index of refraction. Using β/<μe(λ)> = 9.7 x 10-5 kg/m2 in Eq. 14, data reduced by
Method II correlate within 15 % over four decades. Several possible reasons can be
postulated to explain the discrepancies, one being the severe uncertainty in the complex
index of refraction and another being that optical properties actually vary by a factor of
two or more as a function of particle morphology and H/C ratio. Independent
measurements to establish an absolute calibration of <Cm>, therefore, are highly
desirable.
An improvement on the hardware has been described which may simplify the
alignment and allow for the housing of all critical equipment in a single location (e.g., the
detectors may be housed together with the light source and optics). Other new features of
the hardware, beyond the introduction of an accurate beam reflection device, involve the
use of a single photodiode thus improving the accuracy of the measurements. Attention
to the problem of beam steering will undoubtedly lead to further hardware modifications.
The most salient new calculational feature is the use of an empirical material
constant, namely, the average absorption parameter <μe(λ)> . This parameter exhibits a
relatively constant numerical value in spite of the extreme variety of microscopic
mechanisms that characterize soot suspensions. Microscopic details are complicated in
part because soot particles range between 5 and 300 nm in equivalent diameter. This is
the realm of free clusters, also known as microparticles, which exhibit much unusual
behavior and which are the source of present research activity (24, 25). Also, within soot
can be found fullerenes, or buckyballs (C60), allotropes of carbon which differ from
diamond and graphite in crystalline structure.
The software may for certain applications be further simplified by the use of a 2nd
degree polynomial in Eq. 10. This would require transmission measurements at only two
separate wavelengths (the number of wavelengths must match the degree of the
polynomial). Matrix inversion techniques are routinely carried by microprocessor
software (MATLAB, for example). Also, for improved accuracy, wavelengths in the
near- to mid-infrared would be desirable because here there is greater insensitivity of the
Mie theory to particle size and shape. But the reference wavelength should be in the
violet to keep uncertainties in “k” at a minimum. Optical phase conjugation (OPC) is yet
8
another reflection-technique that may be applied perhaps with a BaTiO3 crystal replacing
the retroreflector, but here thermal blooming effects must be compensated for (26), see
Fig.5 (not shown is a needed polarizer inserted before the OPC-device).
REFERENCES
1. Cashdollar, K. L., Lee, C. K., and Singer, J. M., “Three-wavelength Light
Transmission Technique to Measure Smoke Particle Size Concentration”, Applied
Optics, 18, pp. 1763, 1 June 1979.
2. Few, J. D., Lewis, J. W. L., and Hornkohl, J. O., “Optical
Measurements of Turbine Exhaust Particulates”, Report No. NAPC-PE-
221C, July 1991.

3. Few, J. D. and Hornkohl, J. O., “Measurement and Prediction of Jet


Engine Particulate Effluents”, Report No. NAPC-PE-195C, February
1990.

4. Few, J. D., Lewis, J. W. L., and Hornkohl, J. O., “Optical Particle Sizing
Measurements in the Exhaust of an AGT 1500 Combustor”, Report
NAVAIRWARCEN AC DIV-PE-235C, July 1992.

5. Hirleman, E. D., “Nonintrusive Laser-Based Particle Diagnostics”,


Combustion Diagnostics by Nonintrusive Methods, McCay and Roux
Eds., AIAA Progress in Astronautics, 92, 1984.

6. Kattawar, G. W. and Plaas, G. N., “Electromagnetic Scattering from


Absorbing Spheres”, Applied Optics, 8, pp. 1377, August 1967.

7. Stull, V. R. and Plaas, G. N., “Emissivity of Dispersed Carbon


Particles”, J. Opt. Soc. of Am., 50, pp.121, Feb. 1960.

8. Wade, R. A., Sirvatham, Y. R., and Gore, J. P., “Soot Volume Fraction
and Temperature Properties of High Liquid Loading Spray Flames”, The
Combustion Institute (Eastern US Section), Chemical and Physical
Processes in Combustion, 1993 Technical Meeting, Princeton, NJ, Oct.
25-27, 1993.

9. Baron, P. A. and Willeke, K., “Aerosol Fundamentals”, Aerosol


Measurements, Willeke and Baron Eds., Van Nostrand Reinhold, New
York, 1993.

10. Dobbins, R. A., Santoro, R. J., and Semerjian, H. G., “Interpretation


of Optical Measurements of Soot Flames”, Combustion Diagnostics by
Nonintrusive Methods, McCay and Roux Eds., AIAA Progress in
Astronautics, 92, 1984.

11. Powell, E. A. and Zinn, B. T., “In Situ Measurements of the


Complex Refractive Index of Combustion Generated Particulates”,
9
Combustion Diagnostics by Nonintrusive Methods, McCay and Roux
Eds., AIAA Progress in Astronautics, 92, 1984.

12. Van de Hulst, H. C., Light Scattering by Small Particles, Dover


Publication Inc., New York, 1981.

13. Williams, J. R., et al, “Thermal Radiation Absorption by Particle


Seeded Gases”, AIAA J. Spacecraft and Rockets, 8, pp. 339, April 1971.

14. Shenoy, A. S., “The Attenuation of Radiant Energy in Hot Seeded


Hydrogen”, PhD Dissertation, Georgia Institute of Technology, School
of Nuclear Engineering (164 pages), May 1969,

15. Krascella, N. L., “Theoretical Investigation of the Absorption and


Scattering Characteristics of Small Particles”, United Aircraft
Corporation Report NASA CR-210, April 1965.
16. Marteney, P. J., “Experimental Investigation of the Opacity of Small
Particles”, United Aircraft Corporation Report NASA CR-211, April 1965.

17. Holman, J. P., Experimental Methods for Engineers, Ch. 3, Vth Ed.,
McGraw-Hill, New York, 1989.

18. Bramer, J. R. and Netzer, D. W., “An Investigation of the


Effectiveness of Smoke Suppressant Fuel Additives for Turbojet
Applications”, Naval Postgraduate School Report, NPS67-82-13, March
1982.

19. Dubeau, R. W., et al, “An Investigation of the Effects of Fuel


Composition on Combustion Characteristics in a T-63 Combustor”,
Naval Postgraduate School Report, NPS67-85-004, March 1985.

20. Bennett, J. S., Jway, C. H., Urich, D. J., and Netzer, D. W., “Gas
Turbine Combustor and Engine Augmentor Tube Sooting
Characteristics”, Naval Postgraduate School Report, NPS67-86-004,
December 1986.

21. Young, M. F., Grafton, T. A., Conner, H., and Netzer, D. W.,
“Measurements of Gas Turbine Combustor and Engine Augmentor
Tube Sooting Characteristics”, Naval Postgraduate School Report,
NPS67-88-002, July 1988.

22. Retroreflectors, Melles-Griot Optics Guide 5, pps. 1013-1015 (1717


Kettring Str. Irvine, CA, 92714).

23. Sugimoto, N. and Bleier, Z., “Optical Characteristics of Cube-


Corner Retroreflectors and Retroreflector Arrays for Long Path
Absorption Measurements”, PLX inc. Report (40 west Jefryn Blvd., Deer
Park, N. Y., 11729).
10
24. Hayashi, C., “Ultrafine Particles “, Physics Today, 40, pp. 44, Dec.
1987.

25. Huffman, D., “Solid C60”, Physics Today, 44, pp. 22, Nov. 1991.

26. Glaros, G. E., “Soot Particle Density Determination from a Laser


Extinction Multipass Technique”, MSAE Thesis, Naval Postgraduate
School, September 1994.

11
12
Figure 3. Comparison (45° line) of measured soot concentrations
using conventional (Method I) and proposed here (Method
II) techniques.

13
Figure 4. Schematic of apparatus for transmission measurements
using a conventional retroreflector-mirror unit.

Figure 5. Schematic of apparatus for transmission measurements


using an OPC- mirror unit.

14

You might also like