Enviromental Effects Space Shuttle Solid Rocket Motor Exhaust Plumes
Enviromental Effects Space Shuttle Solid Rocket Motor Exhaust Plumes
Enviromental Effects Space Shuttle Solid Rocket Motor Exhaust Plumes
ENVIRONMENTAL EFFECTS
OF SPACE SHUTTLE SOLID ROCKET
MOTOR EXHAUST PLUMES
By
> :* ••
i I
| 'R» t
Prepared under Contract No. NAS1-14271
By
for
f\JASA
National Aeronautics and
Space Administration
>, ->
•/.95^y
BIBLIOGRAPHIC DATA 1. Report No. S.N^ecipient's Accession No.
SHEET
4. Title and Subtitle '5. Report Date
July 1976
ENVIRONMENTAL EFFECTS OF SPACE SHUTTLE SOLID ROCKET 6.
MOTOR EXHAUST PLUMES
7. Author(s) 8- Performing Organization Rept.
No
BaoChuan Hwang and Harold S. Pergament - AeroChem TP-343
9. Performing Organization Name and Address 10. Ptoject/Task/Work Unit No.
AeroChem Research Laboratories, Inc.
P.O. Box 12 11. Contract/Grant No.
Princeton, New Jersey 08540 NAS1-14271
12. Sponsoring Organization Name and Address 13. Type of Report & Period
Covered
National Aeronautics andJSpace Administration Final Report
T.T.-,„I-,•
Washington, „„<-,,,, r>'r>
DC Or>£/.£ '<"' . }', i "X •' i" •'•• s"*-- ,-•. •
20546 14.
16. Abstractsj-he deposition of NOX and HC1 in the stratosphere from the Space Shuttle solid
rocket motors (SRM) and exhaust plume is discussed. A detailed comparison between stratc
spheric deposition rates using the baseline SRM propellant and an 'alternate' propellant
which replaces ammonium perchlorate by ammonium nitrate, shows the total NOX deposition
rate to be approximately the same for each propellant, ranging from ^0.018 kg/m at 15 .
km to fy 0.0036 kg/m at 30 km. For both propellants the ratio of the deposition rates of
NOX to total chlorine-containing species is negligibly small.
Rocket exhaust ground cloud transport processes in the troposphere are also ex-
amined. A brief critique of the NASA/MSFC Multilayer Diffusion Models (presently used
for predicting pollutant deposition in the troposphere) is presented, and some detailed
cloud rise calculations are compared with data for Titan III C launches. The results
show that, when launch time meteorological data are used as input, the model can reason-
ably predict measured cloud stabilization heights.
17. Key Words and Document Analysis. 17a. Descriptors
Space Shuttle
Atmospheric Interaction Plume Program
NASA/MSFC Multilayer Diffusion Model
Page
1. SUMMARY 1
II . INTRODUCTION 2
VII. REFERENCES 21
111
LIST OF TABLES
Table Page
I PROPELLANT COMPOSITION 24
LIST OF FIGURES
Figure
4 SRM nozzle exit plane gas velocity and flow angle distributions 33
iv
Figure Page
EXHAUST PLUMES
I. SUMMARY
This report discusses the effects of the Space Shuttle solid rocket
motor exhaust plumes on stratospheric NOX and HC1 deposition rates. The
stratospheric exhaust plume predictions were made using both the AeroChem
Atmospheric Interaction Plume Program (AIPP) and Low Altitude Plume Program
(LAPP) . The former code treats the coupled effects of nonequilibrium chem-
istry, gas/particle nonequilibrium and plume shocks (including the Mach
disc) ; the latter code specifically treats the influence of nonequilibrium
chemistry in the mixing/afterburning regions and is used to determine after-
burning effects at the lower stratospheric altitudes. Detailed comparisons
are made between stratospheric deposition rates using the baseline SRM pro-
pellant and an 'alternate' propellant which replaces ammonium perchlorate
by ammonium nitrate in order to reduce the total amount of HC1 in the ex-
haust. It was found that the total NOX deposition rate is approximately the
same for each propellant, ranging from = 0.018 kg/m at 15 km to =» 0.0036
kg/m at 30 km. For both propellants the ratio of NOX to total chlorine-
containing species deposition rates is negligibly small, on the order of
10~2 for the baseline propellant and ranging from 10"1 to 10~2 for the
alternate propellant. It is concluded that the alternate propellant, al-
though it contains a large amount of ammonium nitrate, will not increase
stratospheric NC^ deposition rates
The results of detailed calculations, using the AIPP code, are given
for the SRM plume at 30 km altitude and demonstrate the unique capabilities
of this new code.
This section discusses the influence of altitude on NOX and HC1* deposition
rates in the stratosphere for both the baseline and alternate A propellants. ** A
comparison between the propellant formulations (Table I) shows that, in the alter-
nate propellant, ammonium nitrate has replaced most of the ammonium perchlorate
which should result in substantially decreased HC1 concentrations at the nozzle
exit plane. Interestingly enough, although we might expect that the large amount
of ammonium nitrate might result in significantly larger amounts of NOX produced
in the chamber, equilibrium calculations (Table II) show that this is not the
case. In fact, the NO mole fraction is observed to be about a factor of 50 less
than that for the baseline propellant, primarily because of the substantially
decreased chamber temperature. However, the question arises of whether, because
of the ammonium nitrate in the propellant,''' equilibrium values of NOX will in
fact be achieved in the chamber. (It is well known5'6 that in laboratory flames
and furnaces, above-equilibrium levels of NOX are reached.) This question is
addressed below.
The model used for these calculations assumed that all exhaust products,
except the nitrogen-containing species, reached their equilibrium levels at the
computed adiabatic flame temperature. For conservatism, we assumed that 25% of
the N in the propellant was initially converted to NOf (note that at equilibrium
only 0.1% of the N in the propellant is converted to NO), the remaining N going
to N2. A constant pressure, constant temperature, nonequilibrium chemistry
calculation was then made in which the nitrogen-containing species concentrations
were determined by solving the species conservation equations using the initial
conditions given in Table III and the reaction mechanism in Table IV. Figure 1
shows the computed NO mole fraction (XJJQ) as a function of time (labeled FIRST
ASSUMPTION). It is seen that XNO approaches its equilibrium value in about
5 ms . In order to determine whether equilibrium will be achieved in the cham-
ber the above reaction time must be compared with a characteristic flow time.
For an average gas velocity in the chamber of 91.5 m/s§ the time to traverse
* Almost all the chlorine in the exhaust plume is tied up as HC1, although a
small percentage appears as Cl and C12. When the term HC1 is used in the text
it refers to the total chlorine-containing species.
§ This was computed from the SRM mass flow, the density of the combustion pro-
ducts and the approximate cross sectional area of the chamber.
the approximately 61 m length of the chamber is about 700 ms. Therefore we
expect that NOx equilibrium will be achieved under these conditions. In order
to get an approximate upper limit to X^Q in the chamber the rate coefficients
were arbitrarily decreased to their estimated lower bounds (as noted in Table IV),
These results, labeled in figure 1 as, MOST CONSERVATIVE, show that within about
20 ms (1.8 m) equilibrium is reached in the chamber.
(m
NO }T
OPEF = (1)
(m }
NO ex
x
where m,Trt is the mass flow of NOx, subscript T signifies the total amount leav-
NOx
ing the plume and subscript ex signifies the amount leaving the nozzle. For the
purpose of determining the influence of altitude on NOX deposition it is reason-
able to assume that the NOX produced by shocks and afterburning are independent
quantities and can be separately calculated. Then the predicted NOX deposition
rate obtained by adding the NOx produced by shocks and afterburning individually
will give a conservative upper bound to the real NOx deposition. Defining total
mass flow of NOx separately for shocks and afterburning gives,
(2a)
(2b)
Then, defining individual OPEF's for afterburning and shocks gives
(OPEF) = .X . (3)
SH (m_ )ex
x
I.AFT
(OPEF)
AFT ' -tf±l <4)
j ex
x
where (YJJQ ) and (YJJQ ) are the mass fractions of NOX downstream of the
Mach disc and at the exit plane respectively and m^ and mex are the mass
flows passing through the Mach disc and leaving the nozzle, respectively. The
use of equation 6 alone to establish the enhancement of NOX due to shocks im-
plies that the major increase is due to the first Mach disc. The other shocks
(exhaust gas and air shocks and subsequent Mach discs) are assumed to contri-
bute a negligible amount to the total plume NOX enhancement. This assumption
is quite adequate for the plumes of interest (see, e.g. ref. 2).
5
4
(OPEF)AFI - 1 +"£•
pu Y -
rdr
T (8)
NO ex ex
x
where p and u are the gas density and velocity respectively and r is the radial
distance from the axis. From equation 8 it is observed that the influence of
afterburning on plume NOx enhancement roust be determined via a radial integra-
tion of NOX profiles through the plume.
Table VI gives the results of the OPEF calculations showing the relative
effects of shocks and afterburning for both the baseline and alternate propel-
lants. As expected, because of increased reaction rates resulting in greater
heat release in the plume, the effects of afterburning are greater at 15 km
than at 30 km. The enhancement for the alternate propellant is greater than
that for the baseline primarily because the mass flow of NOX leaving the nozzle
is very much less than that of the baseline propellant (see exit plane composi-
tion in Table V).
Since both NOX and HC1 participate in the ozone destruction cycle (NOX
less efficiently than HC1), of primary interest is the ratio of NOX to HC1
deposition rates in the stratosphere. Since there are no major sources or
sinks for the total amount of chlorine contained in the exhaust products, the
mass flow of HC1 leaving the plume is approximately equal to that leaving the
nozzle.. Table VIII gives both the total NOX deposition rate (in kg/m for use
in statospheric diffusion models which use pollutant deposition as a source
term) and the ratio of NOX to HC1 deposition. These results show that the
amount of NOX deposited in the stratosphere is negligibly small compared to
the HC1 deposited,* the NOX/HC1 ratio ranging from values on the order of
10"1 to the order of 10~3.
* Note that this is the case even using the upper bound calculation described
above.
IV.. COMPUTATION OF EXHAUST PLUME PROPERTIES
This section discusses the use of and results from a new computational
tool developed by AeroChem to compute rocket exhaust plume properties, taking
into account the coupled effects of nonequilibrium chemistry, turbulent mixing,
gas/particle dynamic and thermal nonequilibrium and plume shocks. This code,
the Atmospheric Interaction Plume Program (AIPP)8»9 was written with support
from the Air Force Rocket Propulsion Laboratory and has been applied here to
the determination of stratospheric SRM plume properties.
The AIPP code was delivered to the NASA/Langley Research Center and
demonstrated to operate on the CDC 6600 computer facility. Complete documenta-
tion of the code (both user oriented and program oriented) is given in AeroChem
TP-302a and TP-328, which have been updated and submitted as the Computer Pro-
gramming Documentation, together with a source language deck of the code and
test case input and output.
Use of the code requires that detailed nozzle exit plane gas and particle
properties be specified as initial conditions together with the vehicle flight
velocity and altitude, a suitable chemical reaction mechanism and rate coeffi-
cients. The proper nozzle exit plane conditions can be obtained via the
AeroChem fully-coupled nozzle code (FULLNOZ)10 which outputs exit plane gas and
particle properties in a form directly usable as input to the AIPP code.
The AIPP code has been used to calculate complete SRM exhaust plume
properties at 30 km for a flight Mach number of 3.8. Calculations were made
for an equivalent single nozzle (with a mass flow equal to the total mass flow
of the two SRM's). Distributions of gas and particle properties at the nozzle
exit plane (computed via FULLNOZ) are given in figures 3 to 5. The rather
large decrease in gas temperature toward the nozzle lip occurs at the limiting
particle streamline (i.e. there are no particles at radii greater than the
radius at which the limiting particle streamline occurs). Table IX gives the
chemical reaction mechanism and rate coefficients used for the plume calculations.
The shock structure in the plume is given in figure 7, together with the computed
mixing/afterburning region and the dividing streamline (DSL) in the subsonic/
supersonic (SS) region which separates the exhaust gas passing through the Mach
disc from the exhaust gas which has passed through both the exhaust gas shock and
reflected shock. The DSL in the SS region is seen to resemble a convergent
nozzle until the point where the initially subsonic flow turns supersonic. Fig-
ures 7 and 8 give the centerline distributions of temperature, Mach number, pres-
sure and NOX mole fraction. As expected, the pressure drops well below the
ambient pressure just upstream of the Mach disc and then increases to about a
factor of 2 greater than ambient after passing through the normal shock. The
subsequent pressure distribution is observed to rapidly approach the ambient
level.*
* In reference 9 it is pointed out that the AIPP code does not handle multiple
normal shocks downstream of the Mach disc.
The temperature immediately downstream of the Mach disc increases to
approximately the combustion chamber value and then, because the flow is
far out of nonequilibrium, rapidly dissociates with a large temperature
decrease. The centerline mole fraction of NOX (fig. 8) increases to a peak
value that, because of the rapid dissociation, is less than the equilibrium
value at the post-shock pressure and temperature.
The radial plots of X^oX (fig- 9) are indicative of the total NO,, mass
^oo .
flow, i.e., mwn
wwx « jo
\ pu XNn
wux rdr, at any axial location. This mass flow is
then divided by the exit plane NOX mass flow to get the Local Plume Enhancement
Factor (LPEE). Finally, the asymptotic value of the LPEF is the OPEF, which
gives the total NOX deposition rate. A comparison between the OPEF computed
via the AIPP code and that computed from the 'equivalent AIPP' calculations of
reference 1 shows that the results are similar.
1. Heterogeneous Chemistry
RI +• R2 + S -»- P + S
d[R2]
dt riET
where [Ri] and [R2] are the number densities of RI and R2 respectively and [S]
is the particle number density. If we assume that the particle diameters are
small compared to the gas mean free path, then the arrival rate of molecules
at the surface is simply the random flux of molecules across a surface, deter-
mined from kinetic theory. Using this assumption, the rate equation becomes,
where rp. is the radius of the ith particle group, mR is the mass of Rx, k
is the Boltzmann constant and T is gas temperature.
Utilizing equations 9 and 10 and defining a sticking probability, ns» to
account for the fact that not all molecules will stick to the particle surface
gives the following expression for
In a practical sense, all radicals in the flow can recombine at particle sur-
faces and a heterogeneous reaction mechanism (Table X) must be added to the gas
phase reactions given in Table IX.
The formal method of incorporating a heterogeneous reaction scheme
into AIPP is given below:
dt P\ m L
where mL is the mass of the condensable species.
a. Particle Growth - When the condensate particles are small, e.g.,
formed by heterogeneous nucleation via ions, the condensation rate and therefore
the rate of particle growth is controlled via chemical kinetic processes. When
the particles are sufficiently large that the Knudson number (ratio of mean free
path to particle diameter) in the flow falls below unity, the particle growth
(condensation) process is controlled by diffusion of the condensing species
through the boundary layer surrounding the droplet. Basically, the problem
reduces to one of determining the rate of heat transfer from and mass transfer
to a spherical droplet within a surrounding gas stream — exactly the opposite of
the 'evaporating* droplet problem.
' TE ' T
yc
Pr = - (16)
Sc
'
10
From similarity it can be shown that,
Employing the above formulation reduces the basic problem of determining the depo-
sition of a condensing vapor on a particle to that of determining the functions
fi and f2. Considerable theoretical and experimental work has been done on this
fundamental problem covering a wide range of Reynolds numbers (see, e.g. ref. 14).
For 102 < Re < 10s, a laminar boundary probably exists over the front face of a
spherical particle; the flow behind the particle is the 'usual' type of wake flow.
At higher Re, transition to turbulent flow will occur (the precise value depend-
ing in part on the 'freestream' turbulence level). A reasonable semi-empirical
expression for the condensation rate can be obtained from13
Although the particle diameter is increasing with time the quasi-steady approxi-
mation, which assumes the time-dependent rate of change of surface area, 8A/3t,
is much smaller than that computed from the condensation rate m, may safely be
employed. Using the local flow conditions and initial droplet diameter, d^, a
new droplet diameter df at the end of a given integration step, At, can be
computed from
•where p is the particle density. In this step by step manner the increase of
droplet diameter can be determined throughout the plume. In the AIPP code9 this
analysis should be added to subroutine PART.
11
--»>•
3
««>
where n is the number of particles per cm . The classic approach to determining
the kinetics of coagulation is the theory of Smoluchowski16 who proposed that
particles collide due to Brownian motion and that the particles stick to each ~
other in a certain fraction of those collisions. For all particles of the same
size the Smoluchowski expression reduces to
(25)
where R is the universal gas constant, N is Avogadro's number and s is the ratio
of the sphere of influence to the radius of the particle (which, if greater than
unity, assumes that particles will coagulate even before touching). A routine to
determine the coagulation of particles can readily be added to subroutine PART
in the AIPP code.9
* Titan III is used for the environmental effects study because it uses the same
propellant as the SRM and is about 1/2 the thrust. The interpretation of these
data is therefore very significant with regard to the SRM ground cloud study.
12
the errors in predicting the concentration and dosage fields (for different
weather conditions) resulting from transport of the Shuttle ground cloud ex-
haust products as they move downwind in the earth's surface mixing layer. In
this section it is assumed that the reader has a general familiarity with both
rocket exhaust ground cloud dispersion problems and with the MDM code. Addi-
tional information on these topics can be found in references 21 and 22.
The work reported below is the first part of the overall program to validate
the MDM code in the manner described above; it will be continued over the next
year under Contract NAS1-14504.
The basic code includes two programs which describe two different physical
processes; the first describes the buoyant force dominated cloud rise while the
second describes the diffusion processes controlled by atmospheric turbulence.
1. Cloud Rise
The hot exhaust products formed during rocket launches rise due to
buoyancy. The exhaust cloud of pollutants mixes with ambient air because of
the velocity shear between the cloud gases and the air. This entrainment of
cool air has a critical effect on cloud rise since it reduces the buoyant force.
In stable and neutral atmospheres, the cloud will rise to a stabilization height
where it is in thermal equilibrium with the ambient environment, following which
atmospheric turbulence will dominate the subsequent transport process, i.e., fur-
ther motions of the cloud will be controlled by both the wind velocity and turbu-
lent diffusion.
The NASA/MSFC models for predicting cloud rise are based on the work
of Briggs23'2* who developed formulae for stack plumes. The NASA/MSFC MDM in-
corporates two types of cloud rise models: (1) instantaneous cloud rise models
which assume spherical entrainment (designed to apply to normal launches of
solid-fueled rocket vehicles, e.g., Space Shuttle, Titan III, Minutemen II, etc.)
and (2) continuous cloud rise models which assume cylindrical entrainment (used
for. the cloud rise from vehicles with liquid-fueled first stages). For a vehicle
using both liquid and solid fuel engines, e.g., the Delta-Thor, the authors of
the code17 recommend an average of the rise predicted using the instantaneous
and continuous cloud rise models.
13
that the plume grows linearly* (the entrainment assumption), (3) the equivalency
between the mean horizontal speed of the plume in the bent over stage and the
ambient wind speed, and (4) a constant stability parameter S. The entrainment
constant for the linear growth of the radius of the rising cloud is still "a
controversial number.25 The value of 0.64 used in the NASA/MSFC models is based
on a best fit for a bulk of empirical measurements on stack plumes. Certainly,
a check of this value against the rocket exhaust cloud is needed since the
entrainment constant may not be the same for rising plumes which have been gen-
erated by different means.26
In addition to incorporating the Briggs model for the path of the mass
center of the rising cloud, the NASA/MSFC cloud rise models also provide calcula-
tions of the dimensions and spatial positions of the ground cloud at stabilization
and the distribution of exhaust products which are required input in the NASA/MSFC
diffusion models. At present only the input information from MDM models 3 and 4
(see Section V.A.2 below) are given in the cloud rise models. It is assumed that
the pollutant concentration distribution of the stabilized cloud is Gaussian. The
center of the Gaussian distribution is at the cloud stabilization height, calculated
using Briggs' formula. Assuming that the pollutant concentration at a distance of
one radius, r, from the cloud stabilization height, 2^, is 10% of the concentration
at Zjjj, the standard deviation of the Gaussian distribution can be given as r/2.15.
From the entrainment assumption in Briggs1 formula r = Y-Z = 0.64Z, where y is the
entrainment constant defined above. Since the portion of the cloud above the
mixing layer height (also referred to as the inversion height) is dominated by a
convective mechanism, only the portion of the cloud beneath the inversion height
is considered in the NASA models, i.e., the portion of the Gaussian distributed
cloud below the inversion height is used for the diffusion calculation. For
model 3, it is assumed that the portion of the cloud below the inversion height
is completely contained within an elliptical volume, the center of which is lo-
cated at a so-called effective height. Determination of this effective height is
illustrated in figure 10, in which H is the effective height, the shaded area is
the effective ground cloud geometry for model 3 and 2^ is the cloud stabilization
height. In model 4, the stabilized ground cloud in the mixing layer is assumed to
be contained in a number of cylindrical volume sources extending from the ground
to the top -of the inversion height. The distribution of exhaust product concen-
trations in each sublayer for model 4 is assumed to be Gaussian in the alongwind
and lateral directions. The total amount of product in each sublayer is given
as the integration of the corresponding portion of the vertically ascending
distribution cloud. Thus the predicted downwind concentrations will be very
strongly dependent upon the value chosen for the inversion height.
14
2. Atmospheric Diffusion Models
The MDM is derived basically from Gaussian plume theory coupled with
the Cramer diffusion coefficients. The program was originally developed for
hazard estimates for the U.S. Army27 and then adopted by Dumbauld et al22 for
NASA operation. Fundamentally, there are four models included in the program:
the cylindrical distribution model 1, the static-plume model 2, the ellipsoidal
distribution model 3 and the multi-layered distribution model 4. In addition,
three options can be utilized within each model: precipitation scavenging, gra-
vitational settling and surface absorption.
All models in the MDM code assume no flux of material across the
physical layer boundaries. Any flux of material (which would occur if the damp-
ing factor for boundary absorption is not set equal to 1) is assumed to be
instantaneously depleted. Thus the material across a layer boundary will not
be accounted for in the diffusion process in the next layer when multilayer
distribution is calculated simultaneously.
3. Program Output
15
a. Concentration and Dosage - The total time integrated concentration
at a receptor position (defined as dosage) is not directly integrated from the
instantaneous concentration distribution. This is probably because it is not
analytically integrable nor is it easily integrated by numerical techniques.
Therefore it is important to closely examine how the dosage calculation is for-
mulated within the code in order to clearly identify the implicit assumption
associated with its use. In fact, the formulation can be obtained by either
of two possible explanations. The first utilizes the classical assumption that
diffusion in the alongwind direction is negligible in comparison to the gross
transport by the mean wind.30 In general, however, this assumption is not true
for modeling the diffusion of self-contained rocket clouds.31 Figure 11 demon-
strates the differences in the dosage formulation between using and not using
the above assumption. The lower cigar-like figure (lib) (in three dimensions
it looks like a pancake) shows the model used in the MDM for the dosage cal-
culations, which are obtained by neglecting diffusion in the X direction. The
second method of interpretation which will give the identical dosage formulation
is to assume that the effective instantaneous cloud during the time the cloud
passes over an observer has a constant distribution which is identical to the
distribution in the cloud at TQ. TQ is the time at which the cloud centroid
is at the same downwind position, x, as the receptor. The latter method is
also used for the formulations in MDM of the time mean alongwind concentration
which is the average partial dosage from time To-TA/2 to time TQ + TA/2 over
the average time TA. Figure 12 illustrates the second assumption used for the
dosage formulation. This assumption, however, is not correct because the cloud
will diffuse and cannot remain uniform in turbulent flow fields. The assumption
will result in a high predicted value in the near field of the launch pad. It
is apparent that the time mean alongwind concentration multiplied by the time
T. is equivalent to the dosage when T is larger than the cloud passage time.
A A
4. Diffusion Coefficients
The MDM code is based in large part on the Cramer diffusion coeffi-
cients,32-3A which are used to model the physics of atmospheric turbulent trans-
port. Cramer assumed a power law for the diffusion coefficients which, mathemat-
ically, are the standard deviation of the pollutant particle dispersion
displacement,
16
oy = aQ x P (26)
Two sets of experimental data, from the Prairie Grass35 and Round
36
Hill diffusion studies, were used to support Cramer's diffusion coefficients.
However it should be mentioned that both the Prairie Grass and Round Hill diffu-
sion experiments were carried out in planetary surface layers, which are usually
on the order of only tens of meters. Thus, it is questionable whether the Cramer
coefficients are suitable for investigating rocket ground clouds in the atmo-
spheric mixing layer with heights on the order of hundreds to a thousand meters.
At present the standard deviations ae and a^ of azimuth angle.and elevated angle
used in the MDM are calculated from an ad hoc formula22 based on the measured
a. only at a reference height Z near the ground. These values need to be
verified both theoretically and experimentally.
5. Code Operation
The total program contains about 6000 Fortran statements and 44 sub-
routines, most of which are for output purposes. Our experience has shown that
modifications to the code will be quite difficult and could cause operational
problems.
Both the cloud rise (preprocessor) program and the main MDM program
were written in Fortran IV, originally for a Univac computer. Since the pres-
ent operating computer system at NASA/Langley is a CDC 6600 and since there are
some differences in Fortran IV software among the various computer systems, a
conversion of the code to the CDC machine from the Univac system was needed.
This has been accomplished on version 5 of the code and several test cases have
been run successfuly. using different models. The CDC version of the NASA/MSFC
MDM program can be obtained from either NASA/Langley or AeroChem upon request.
1. Meteorological Conditions
17
velocity, temperature, wind direction and relative humidity are shown in fig-
ure 13. From these measured meteorological data collected near the launch
complex at the time of launch, the inversion height appears to be at about
2203 m from the ground. In fact, the consistent sharp changes of humidity,
temperature and wind velocity near 2203 m demonstrate the existence of the inver-
sion layer. It is evident that at launch time the atmosphere was in a slightly
stably stratified condition. Using Tabata's formula,37 in which the effect of
relative humidity is considered empirically, the potential temperature gradient
turns out to be about 0.011 degrees K per meter.
Using the meteorological data mentioned above and the NASA/MSFC cloud
rise models, the cloud was calculated to be stabilized at a height of 1527 m
and 1664 m away from the launch pad. The total time to reach the stabilization
height was about 8 min. 45 sec. The temporal history is estimated from the cloud
rise formula for stable conditions in the MSFC instantaneous cloud rise models.
For a Titan III C normal launch vehicle, the expression is,
where t(sec) is the time for the exhaust cloud to rise to an altitude Z (meter).
Pg (g/m3) is the air density at the surface and Az<|> is the potential temperature
gradient.
The stability parameter S is,
3. Cloud Width and Cloud Shape Calculations and Comparison with Data
The growing history of the cloud crosswind width during cloud rise is
calculated by using the following formula, given by the NASA/MSFC cloud rise
models,
where r is the cloud radius at any height Z less than the cloud stabilization
height Zm and y is the entrainment constant (0.64).
18
The cloud crosswind width history is shown in figure 15 in which the
shaded area is the envelope of cloud widths from the NASA/Langley measurements
for four launches.* The predicted cloud crosswind width'in the initial stage
apparently grows much faster than the average measured values but goes into the
envelope near the cloud stabilization height.
The four launches are February 11, 1974, May 30, 1974, December 10, 1974
and March 14, 1976.
19
VI. CONCLUDING REMARKS
2. The value selected for the inversion height and the entrainment
constant y may strongly influence the downwind deposition prediction.
A detailed investigation of these two parameters is needed.
20
VII. . REFERENCES
2. Pergament, H.S., Thorpe, R.D. and Hwang, B., "NO Deposited in the Stratosphere
by the Space Shuttle Solid Rocket Motors," Final Summary Report, Phase II,
AeroChem TP-333, NASA CR-144928, December 1975.
3. Thorpe, R.D., Pergament, H.S. and Hwang, B., "NOX Deposition in the Stratosphere
by the Space Shuttle Solid Rocket Motors," JANNAF 9th Plume Technology Meeting,
C.P1A Publ. 277 (Applied Physics Lab., Johns Hopkins Univ., Silver Spring, April
1976) pp. 317-352.
4. Space Shuttle Environmental Workshop on Stratospheric Effects, JSC-11137,
24-25 March 1976
5. Tokagi, T., Ogasawara, M., Fujii, K., Dazo, M., "A Study on Nitric Oxide
Formation in Turbulent Flames," Fifteenth Symposium (International) on Combustion
(The Combustion Institute, Pittsburgh, 1975) pp. 1051-1059.
7. Mikatarian, R.R., Kau, C.J. and Pergament, H.S., "A Fast Computer Program for
Nonequilibrium Rocket Plume Predictions," Final Report, AeroChem TP-282, AFRPL
TR-72-94, NTIS AD 751 984, August 1972.
8. Pergament, H.S. and Kelly, J.T., "A Fully-Coupled Underexpanded Rocket Plume
Program (The AIPP Code). Part I. Analytical and Numerical Techniques," Final
Report, AeroChem TP-302a, AFRPL-TR-74-59, NTIS AD A006 235, November 1974.
10. Pergament, H.S. and Thorpe, R.D.., "A Computer Code for Fully-Coupled Rocket
Nozzle Flows (FULLNOZ)," AeroChem TP-322, AFOSR-TR-75-1563, NTIS AD A019 538,
April 1975.
11. Pergament, H.S. and' Kau, C.J., "A Computer Code to Predict the Effects of
Electrophilic Liquid Injection on Re-Entry Plasma Sheath Properties," Final
Report, AeroChem TP-308,. AFCRL-TR-74-0074, NTIS AD 782 023, January 1974.
12. Jensen, D.E., "Chamber Non-Equilibrium and Smoke Predictions for a Liquid
Monopropellant Rocket Motor," JANNAF 7th Plume Technology Meeting, CPIA
Publ. No. 234, June 1973.
13. Fuchs, N.A., Evaporation and Droplet Growth in Gaseous Media (Pergamon Press,
London, 1959).
21
14. Johnstone, H.F. and Eades, D.K., Ind. Eng, Chem. 42_,. 2293 (1950),
15. Zebel, G., ''Coagulation of Aerosols," Aerosol Science, C.N. Davies, Ed.
(Academic Press, London, 1966) Ch. II.
17. Dumbauld, R.K. and Bjorklund, J.R., "NASA/MSFC Multilayer Diffusion Models
and Computer Programs - Version 5," NASA CR-2631, 1975.
18. Stephens, J.B., "Atmospheric Diffusion Predictions for the Exhaust Effluents
from the Launch of a Titan 3C, December 13, 1973," NASA TM X-64925, 27
September 1974.
19. Lange, R., "ADPIC, A Three Dimensional Computer Code for the Study of Pollutant
Dispersal and Deposition under Complex Conditions," Lawrence Livermore Lab.,
UCRL-5KL62, 1973.
20. Gregory, G.L., Wornom, D.C., Bendura, R.J. and Wagner, H.S., "Hydrogen Chloride
Measurements from Titan III Launches at the Air Force Eastern Test Range, FL,
1973 thru 1975," NASA TM X-72832, 1976.
21. Slade, D.H., Ed., "Meteorology and Atomic Energy," USAEC-TID-24190, 1968.
22. Dumbauld, R.K. et al, ""NASA/MSFC Multilayer Diffusion Models and Computer
Program for Operational Prediction of Toxic Fuel Hazards," NASA CR-129006,
1973.
23. Briggs, G.A., "Plume Rise: A Critical Survey," Air Resources Atmospheric
Turbulence and Diffusion Lab., TID-25075, 1969.
24. Briggs, G.A., "Some Recent Analyses of Plume Rise Observations," presented
at the Second International Clean Air Congress, Washington, DC, December
6-11, 1970, Paper ME-8E.
25. Briggs, G.A., "Discussion; Chimney Plumes in Neutral and Stable Surroundings,"
Atmos. Env. £, 511-512 (1972).
26. Baker, P.J. and Jacobs, B.E.A., "Pulsed. Emission Chimney," Civil Engineering
Publics Work Rev., 199-200 (1971).
27. Cramer, H.E. and Dumbauld, R.K., "Experimental Designs for Dosage Prediction
in CB Field Tests," Geophys. Corp. of America, GCA 68-17-G, 1968.
28. Corrsin, S., "Limitations of Gradient Transport Models in Random Walks and in
Turbulence," Advances in Geophysics, Vol. 18A (Academic Press, New York, 1974)
pp. 25-60.
22
30. Frenkiel, F.N., "Turbulent Diffusion: Mean Concentration Distribution in a
Flow Field of Homogeneous Turbulence," Adv. Appl Mech, _3, 61-107 (1953).
31. Stephens, J.B. and Stewart, R.B., ''"Rocket Exhaust Effluent Modeling for
Tropospheric Air Quality and Environmental Assessment," Preliminary Report,
NASA, 1975.
32. Cramer, H.E., "A Practical Method for Estimating the Dispersal of Contaminants,"
Proceedings of the First National Conference on Applied Meteorology (Amer.
Meteor. Soc., Hartford, 1957) pp. C-33 to C-35.
34. Cramer, H.E. et.al, "Meteorological Prediction Techniques and Data System,"
Geophysics Corp. of America, GCA-64-3-G, 1964.
35. Barad, M.L., "Project Prairie Grass. A Field Program in Diffusion," Geo-
physics Research Papers No. 59, Vols. I and II, AFCRC-TR-58-235, July 1958.
36. Cramer, H.E., Record, F.A. and Vaughan, H.C., Final Report, AFCRC-TR-58-
239, 1958, p. 70.
37. Tabata, S., "A Simple but Accurate Formula for the Saturation Vapor Pressure
Over Liquid Water," J. Appl. Meteor. 12, 1410 (1973).
23
. TABLE I
PROPELLANT COMPOSITION
Wt% AH°(kcal/mole)
—r
Standard Alternate
TABLE II
Standard Alternate
Pressure, Pa (N/m2) 4.14 x 106 4 .14 x 106
Temperature , K 3400 2690
Mass Fraction A1203 Particles 0.30 0.283
CO 2.49(-l) 2.7K-1)
C02 1.74 (-2) 8.63(-3)
HC1 1.43(-1) 1.92 (-2)
H 3.80(-2) 5, 47 (-3)
H2 2.77(-l) 4.19C-1)
OH 9.1K-3) 2. 84 (-4)
H20 1.52(-1) 8.80(-2)
N2 9. 92 (-2) 1.87C-1)
N 6. 19 (-6) < 1.0 (-9)
NO 7. 05 (-4) 1.60C-5)
0 7.45(-4) 2.45(-6)
02 1.62(-4) < l.OC-9)
Cl 1.30(-2) 1.73C-4)
24
TABLE III
TABLE IV
REDUCED NO,,
A.
REACTION SET
Kf = AT~N exp(B/RT)
B Upper, Lower
N (cal/mole) Bound Bound
N + N +M = N2 + M 5.5 (-30) 1.0 0.0 10 10
0 + N2 = NO + N 1.3 (-10) 0.0 -76000 3 3
N -h 02 ^ NO + 0 2.2 (-11) 0.0 6250 3 3
NO + H i: N + OH 1.5 (-10) 0.0 -47000 5 5
cm-molecule-sec units
multiply Kf by upper bound to get upper limit of rate coefficient
divide K,. by lower bound to get lower limit of rate coefficient
25
TABLE V
Standard Alternate
Pressure, Pa (N/m ) 2
12.21 x 103 1.22 x 10"
Temperature, K 1350 1070
Velocity, m/s 2710 2410
Radius, m 8.23 8.23
Composition, Mole Fraction
CO 2.28C-1) 2.39(-l)
CO2 2.47(-2) 1.18 (-2)
Cl 4. 08 (-3) 5. 25 (-5)
HC1 1.44(-1) 1.85(-2)
H 8. 75 (-3) 1.2K-3)
OH 7. 21 (-5) 2.20(-6)
H2 2.74C-1) 4.03(-1)
H20 1.46C-1) 8. 18 (-2)
N2 9.45(-2) 1.72(-1)
0 2. 09 (-6) 6. 63 (-9)
02 7.76(-7) 9. 62 (-10)
Al203(s) 7. 46 (-2) 7. 21 (-2)
N 9. 95 (-10) 9. 62 (-IP)
NO 1.99 (-4) 4. 36 (-6)
TABLE VI
26
TABLE VII
Nozzle 15 km 30 km
Plume Enhancement
TOTAL
TABLE VIII
NO Deposition Rate
X ...
(kg/m)
15 km 30km
Baseline 0.0207 0.0036
Alternate 0.0279 0.0027
27
TABLE IX
K f =AT~" exp(B/RT)
B
A N (cal/mole)
0 + 0 + M -> : 02 + M 1.0 (-29) 1.0 0.0
0 + H + M ->•OH + M l.OC-29) 1.0 0.0
H + H + M' -»• H2 + M 5.0(-29) 1.0 0.0
H + OH + M -»• H20 + M 2.0(-28) 1.0 0.0
CO + 0 + M ->•C02 + M 1.0(-29) 1.0. -2500
OH + OH ->• H20 + 0 l.O(-ll) 0.0 -780
OH + H2 •* H20 + H 3. 6 (-11) 0.0 -5200
0 + H2 ->• OH + H 2.9(-ll) 0.0 -9460
H + 02 -»• OH + 0 3. 7 (-10) 0.0 -16800
CO + OH -»• C02 + H 9.0(-13) 0.0 -1080
H + Cl + M -»• HC1 + M 5.5(-31) 1.0 0.0
HC1 + OH ->• H20 + Cl 7.2(-12) 0.0 -3250
H + HC1 -»• Cl + H2 8. 8 (-11) 0.0 -4620
OH + Cl •*• HC1 + 0 3.0(-11) 0.0 -5000
0 + N2 -»• NO + N 1.3(-10) 0.0 -76000
N + 02 •*• 0 + NO 2.2(-ll) 0.0 -6250
NO + H -»• N + OH 1.5(-10) 0.0 -47000
N + N + M -»• N2 + M 5.5(-30) 1.0 0.0
28
TABLE X
H + OH + S ->• H20 + S
0 + H+S -»• OH + S
0 + 0 + S -»• 02 + S
H + H + S -»• H2 + S
CO 4- 0 + S -»• C02 + S
H + Cl + S -»• HC1 +. S
Cl + Cl + S -»• C12 + S
N + N + S -J- N2 + S
N + O + S -»• NO + S
29
76-102
TEMPERATURE t K
1500
2000
10-2
z
o
h-
o
a:
u_
LU I0~3
_i
o
FIRST ASSUMPTION
MOST CONSERVATIVE
io-4
EQUILIBRIUM
0 15 20 25 30 35
TIME, milliseconds
30
76-101
I n r
3000 120
TEMPERATURE
LPEF
U.
UJ
Q_
2600 100
ALTERNATE
o
h-
CJ
2200 80
Lul UJ
CK 2
UJ
o
< 1800 60
LU
Q_
X
2 UJ
UJ
h-
UJ
1400 40 ^
ZD
BASELINE _l
Q.
o
1000 20 o
600
20 40 60 80
31
76-107
2200
TEMPERATURE
.8XI05
~\J > ^ * "'''
^~ 2100
I.6XI0 5
2000
CVJ
1900
WI.2XI05
PRESSURE LU
CO 1800
CO
Ld LU
o:
Q.
I0
1700
8XI0 4
1600
1500
0 0.5 1.0 1.5 2.0 2.5
RADIAL DISTANCE FROM AXIS,m
32
76-109
0.25 -VELOCITY
0.20
o
o>
0.15 IO
O
o FLOW ANGLE
I o.io O
O
_l
LiJ
0.05
2.5 2500
u
CO
« 2.3 2300
UJ
10
tr
O
2.2 2200 eg
Q.
O UJ
O
_l
LU
2.1 2100
u
UJ I-
O cc
a: 2,0 2000
£
QL
1.8 1 1800
0.5 1.0 1.5 2.0 2.5
34
76-110
AIR SHOCK
120
MIXING/AFTERBURNING REGION
X
<
REFLECTED SHOCK
o SUBSONIC/SUPERSONIC
(T
UL DIVIDING STREAMLINE
80
UJ
O EXHAUST GAS SHOCK
h-
Q
40
O
cr
0
0 40 80 120 160 200
DISTANCE FROM NOZZLE EXIT PLANE, m
MACH NUMBER
(T
UJ UJ
o: TEMPERATURE CD
ID 4000
IT
UJ
Q_ 3 5
5
UJ
2000
MACH NUMBER
0
40 80 120 160 200
DISTANCE FROM NOZZLE EXIT PLANE,m
1 I I I
PRESSURE
10 -2 2
O
I-
o
Q_
AMBIENT PRESSURE (30km)
UJ
UJ
CO
_l
C/5
o
UJ
a: X
Q_
O
10 10'3
I I I I 10-4
80 160 240
DISTANCE FROM NOZZLE EXIT PLANE,m
Figure 8 SRM exhaust plume centerline pressure and NOX mole fraction
distributions.
37
76-114
10 -3
t
I
O 10-4
J-
o
u_
UJ
_l DISTANCE FROM
o NOZZLE EXIT, m
— — — — 140
x
O
40
10-6
10 20 30
RADIAL DISTANCE FROM AXIS, m
38
76-105
3200
2800 -
2400-
~2000 -
1600-
LU
X 1200-
800-
400-
WIDTH, m
39
76-94
INSTANTANEOUS
CONCENTRATION
ISOPLETH
dxnrdt
40
76-100
RECEPTOR
Uk
76-92
•V-i
2400
2000
1600
1200
UJ
800
400
-9.8°C/km
0
10 15 20 25 3.5 7.0 10.5 2 4 6 8 40 60 80 999
1
' TEMPERATURE, °C WIND DIRECTION, WIND SPEED, m sec' RELATIVE HUMIDITY, percent
rod.
i i i i i 1 I I I
CACULATED
LU STABILIZATION HEIGHT
Q
tj<
Q
O
_J —— — NASA/LANGLEY MEASUREMENTS
O
(ASKANIA TRACKING UNIT)
ENVELOPE
NASA/LANGLEY DATA
I
h-
O
Q
D
o
_J
o
Q
CO
CO
o
<r
o
FROM NASA/MSFC
CLOUD RISE MODEL
2 4 6 8 10
TIME AFTER LAUNCH, minutes
44
76-99
3500 l
INVERSION LAYER AT 2203 m
2000
;
1500
LJ
X
1000
500
0
-1000 0 1000 2000 3OOO 4000
DISTANCE FROM LAUNCHING PAD, m
Ol
O\ 76-96
3000
2500
INVERSION LAYER ASSUMED AT'
2203 m
2000
1500
1000
LJ
X
500
0 I I I I I
0 1000 2000 3000
DISTANCE FROM LAUNCHING PAD, m