Skip to main content

Full text of "Analysis of the flow in a 1-MJ electric-arc shock tunnel"

See other formats


NASA TECHNICAL NOTE 



NO 




NASA TN D -6865 



i ^./ 



m 
o 



■OAN COPY: REr^M I 
AFWL (DOUitii^ ;5 



\y\ 



ANALYSIS OF THE FLOW IN 

A l-MJ ELECTRIC-ARC SHOCK TUNNEL 

by John 0. Keller, Jr., and N. M. Keddy 

Ames Research Center 
Moffett Field, Calif. 94035 



'iri iy/2 



NATIONAL AERONAUTICS AND SPACE ADMINISTRATION • WASHINGTON, D. C. • JUNE 1972 



TECH LIBRARY KAFB, NM 



DiB33aa 



lllllll 



1. Report No. 

NASA TN D-6865 



2. Government Accession No. 



4. Title and Subtitle 

ANALYSIS OF THE FLOW IN A 1-MJ ELECTRIC-ARC SHOCK TUNNEL 



7. Author(s) 

John O. ReUer, Jr., and N. M. Reddy 

9. Performing Organization Name and Address 

NASA— Ames Research Center 
Moffett Field, Calif. 94035 



12. Sponsoring Agency Name and Address 

National Aeronautics and Space Administration 
Washington, D. C. 20546 

15. Supplementary Notes 



3. Recipient's Catalog No. 



5. Report Date 

June 1972 



6. Performing Organization Code 



8. Performing Organization Report No. 
A-4341 



10. Work Unit No. 

117-07-04-15-00-21 



11. Contract or Grant No. 



13. Type of Report and Period Covered 
Technical Note 



14. Sponsoring Agency Code 



16. Abstract 

An investigation has been conducted in the Ames electric-arc-heated shock tunnel to evaluate the performance of the 
facility over a range of shock Mach numbers from 7 to 19. The efficiency of the arc-heated driver is deduced using a new 
form of the shock-tube equation. A theoretical and experimental analysis is made of the tailored-interface condition. The 
free-stream properties in the test section, with nitrogen as the test gas, are evaluated using a method based on stagnation- 
point heat-transfer measurements. 



17. Key Words (Suggested by Author(s) ) 
Shock tunnels 
Nozzle flow 
Arc-discharge heating 
Calibrating 
Test facilities 

19. Security Classif. (of this report) 
Unclassified 



18. Distribution Statement 

Unclassified — Unlimited 



20. Security Classif. (of this page) 
Unclassified 



21. No. of Pages 
28 



22. Price 
3.00 



' For sale by the National Technical Information Service, Springfield, Virginia 22151 



TABLE OF CONTENTS 

Page 

NOTATION V 

SUMMARY 1 

INTRODUCTION 1 

THEORETICAL CONSIDERATIONS 2 

The Shock-Tube Equation 2 

Driver Efficiency 3 

Tailored-Interface Operation 3 

Test Section Flow 5 

EXPERIMENTS 6 

The Test Facility 7 

Driver Performance V 

Reservoir Conditions 8 

Test Section Measurements 9 

The three-probe rake 9 

The survey rake 10 

DISCUSSION ..11 

Nozzle Test Core 12 

Total Enthalpy 12 

Free-Stream Properties 14 

Atom concentration 15 

Velocity 15 

Density 16 

Shock-layer viscosity 16 

Shock-layer Reynolds number 17 

Mach number, Reynolds number, and ylM * IV 

CONCLUDING REMARKS 17 

REFERENCES 19 

TABLES 21 



ui 



■iiiiiiiiiiiiiiiiin^niH^niiiiBinininiiiiii i mill 



NOTATION 

a speed of sound 

A constant in equation (21), ^4 =6X10^° cm^ °K/mole sec; effective area of inviscid nozzle 
flow 

cjj coefficient of specific heat for Pyrex 7740 

d diameter 

E internal energy 

E^ capacitor bank energy 

hj^ heat of dissociation per unit mass 

H total enthalpy 

H^ reference enthalpy (300 J/g) 

/ for two-dimensional flow; 1 for axisymmetric flow 

Kfj coefficient of conductivity of Pyrex 7740 

Le Lewis number 

m molecular weight 

M^ shock Mach number (f/^/fli ) 

p pressure 

Pf pitot pressure 

Pr Prandtl number 

q convective heat-transfer rate 

R gas constant per unit mass of undissociated molecules 

Ryi dimensionless velocity number defined in equation (1 ) 

S entropy 

S^ shock tube number defined in equation (2a) 

T absolute temperature 

U flow velocity relative to tube wall 



C/y shock velocity 

C/3 limiting velocity, 2a4/74 - 1 

V volume of the driver 

Z compressibihty, 1 + a 

a atom concentration 

ap temperature coefficient of resistance of platinum film 

7 ratio of specific heats 

6 * boundary-layer displacement thickness 

e density ratio across a normal shock wave, Pi /p2 

M dynamic viscosity 

p gas density 

p^ density of Pyrex 7740 

T} efficiency of the driver 

^ parameter involving 7, defined in equation ( 1 0) 

6 characteristic dissociation temperature; 1 13,260° K for nitrogen 

Subscripts 

reservoir condition at end of shock tube 

1 quiescent test gas ahead of incident shock wave 

2 test gas between incident shock wave and contact surface 

3 driver gas behind incident contact surface 

4 driver chamber conditions after arc discharge 
e edge of the boundary layer 

eq equilibrium boundary-layer model 

fc fully catalytic 

fm free molecule flow model 

fr frozen boundary-layer model 

vi 



/ initial driver condition before arc discharge 

nc noncatalytic 

T tailored interface 

oo free-stream condition at exit of nozzle 

Superscript 

* nozzle sonic throat 



vn 



ANALYSIS OF THE FLOW IN A 1-MJ ELECTRIC- ARC SHOCK TUNNEL 

John O. Keller, Jr., and N. M. Reddy* 
Ames Research Center 

SUMMARY 



An investigation has been conducted in the Ames electric-arc-heated shock tunnel to evaluate 
the performance of the facility over a range of shock Mach numbers from 7 to 19. The efficiency of 
the arc-heated driver is deduced using a new form of the shock-tube equation. A theoretical and 
experimental analysis is made of the tailored-interface condition. The free-stream properties in the 
test section, with nitrogen as the test gas, are evaluated using a method based on stagnation-point 
heat-transfer measurements. 



INTRODUCTION 



The shock driven wind tunnel has played an important role in the study of hypervelocity 
aerodynamics, in part because of the wide range of flow velocities available and the relative ease 
with which the composition of the test stream can be varied. Relatively long duration test flows 
may be obtained with the tailored-interface mode of operation, as discussed in references 1 and 2. 
Heated hydrogen drivers (ref. 3) and combustion drivers (refs. 4, 5, 6) have been extensively used to 
produce tailored shock Mach numbers in the range of 6 to 9. However, higher shock Mach numbers 
are necessary to dissociate and ionize gases such as nitrogen and carbon monoxide, which are of 
interest in planetary entry studies. To achieve such conditions within the constraints imposed by 
the tailored mode of operation requires driver gas temperatures in excess of 2000° K. Arc-discharge 
heating (refs. 7 and 8) has been used successfully in this application; a wide range of driver 
temperatures is made available by varying driver loading pressure or discharge energy. This attractive 
scheme has been used for the Ames electric-arc shock tunnel (EAST), which has a capacitor storage 
system rated at 1 MJ. 

In the present study, the characteristics of the heated driver and the tailored-interface 
operation of the shock tube (with nitrogen) have been considered, both theoretically and 
experimentally. The major emphasis, however, is on evaluating the properties of the highly 
expanded nitrogen stream in the test section. In this low-density environment, the more 
conventional methods of stream evaluation, which rely on probe measurements of static pressure, 
are inaccurate, largely because of substantial corrections for orifice and probe boundary-layer 
effects. Thus, in this study a number of quantities characterizing the flow in the EAST test section 
have been deduced using a new technique based on stagnation-point heating-rate measurements. 
Other flow properties were obtained from these measurements by application of the sudden-freeze 
concepts of reference 6. The results are compared with numerical solutions of nonequilibrium 
nozzle flow using the method of reference 9. 



*NASA— NAS Research Associate. Presently at Indian Institute of Science, Bangalore — 12, India. 



■ II ■■■iiii^ii III ■■■■■III ■■■■■^■^■■■■■illii 



THEORETICAL CONSIDERATIONS 



The idealized performance of an electric-arc-heated driver and shock tube can be predicted 
easily for any driver gas or test gas mixture using the generalized form of the classical shock-tube 
equation presented in reference 10. This form of the shock-tube equation presents simple 
expressions for the tailored mode of operation and provides an easy method of calculating the 
actual efficiency of the electric-arc-heated driver. This information is the basis for estimating 
reservoir conditions at the nozzle inlet for an arbitrary test gas. These conditions in turn are input 
quantities for calculating the flow properties in the expansion nozzle and test section. The concepts 
of reference 10 are reviewed briefly and applied to the arc-heated driver; comparisons are made of 
predicted and realized reservoir conditions. 



The Shock-Tube Equation 
It is shown in reference 1 that by defining a dimensionless velocity 

Rr,^^ (1) 

the classical shock-tube equation, relating particle velocity behind the incident shock wave to driver 
gas properties before diaphragm opening, can be reduced to a universal form given by 






74/(74-') 



Sn-- ""l, -" (2a) 



where 



^ ^'^1 __^li_ ^ _^24_ / Pxlp^ .2b) 



'" 74 ~ 1 yJlxPjPi 74 " 1 V 74 
and 

R =(7^n(l-e)M, ^20 

" 2(aja^') 

In equations (2) the driver gas is assumed to be perfect, but no assumption as to the thermodynamic 
state of the driven gas is made. (Equations (2) are for a constant-area tube and represent the exact 
solution for perfect gases to within 3 percent at Ms > 4.) 

In an electric-arc-heated facility, the driver gas is heated at constant volume and hence the 
initial density p^ is equal to the final density p^ after heating. Hence 

---'''- (3) 

Also 

:i=^ _i (4) 

Pi Pi "'i 



since in general Tj = T^. With the use of equations (3) and (4), the shock-tube equation can be 
reduced to 

,74/(74-0 




y ^ R (5) 

74 ^n 

Thus, the dimensionless shock-tube number Sn, for constant-volume heating of the driver gas, 
depends only on the initial pressures and molecular weights of the driver and driven gases. 
Equation (5) is convenient to use for estimates of shock-tube performance for any combination of 
gases; in graphical form, the variation of Rn with S^i is a single curve for a given 74 • The value of Rfi 
that corresponds to a particular Sn is unique in the sense that it is independent of the 
thermodynamic state of the gas behind the shock wave. Having R^, we use equation (2c) to relate 
Ms directly to 04 and T4, the speed of sound and temperature in the heated driver gas. Obviously, if 
Ms is measured, then an effective T^ can be determined. 



Driver Efficiency 
In an electric-arc-heated driver with constant volume heating of a perfect gas (ref. 1 1 ) 



and 



Ij^Il) (7, 

v7i m^ T^ 

These driver-heating equations can be combined with shock tube equations (2c) and (5) to obtain 
the efficiency of the driver. With a^/ai derived from measured Ms, and equation (2c) with the 
appropriate value of Rn, equation (6) will yield 17. In the present study, 1? was based on Ms 
measured at the downstream end of the shock tube, and as such includes losses associated with 
shock-wave attenuation. It could be reasoned that this approach results in a lower estimated driver 
efficiency, because it does not account for viscous losses in the shock tube; however, the result does 
give an estimate of overall driver effectiveness. The efficiencies so determined are discussed in a later 
section on driver performance. 



Tailored-Interface Operation 

At the downstream end of the shock tube a reservoir of heated gas is produced by reflection of 
the incident shock wave from the end wall. An optimum condition exists if the downstream wave, 
resulting from the interaction of the reflected shock wave with the contact surface, is a Mach wave. 
When this occurs, the contact surface (interface) is said to be "tailored." Since this secondary 
reflection is vanishingly weak, the reservoir conditions remain steady for a longer period — in 
theory, until the arrival of the main driver expansion. For tailored operation a matching condition 
can be derived for the perfect driver and driven gas case, as shown in reference 12; for example. 



[ 



^1 



74 + ^ 7i -1 
74-1 7i + 1 



M^ » 1 



(8) 



The dimensionless number Rn can be used to write the ratio of internal energies, following 
reference 10, as 



E2 



3 _ T4-I (l-Rn 



274 



Rr 



From equations (8) and (9) the tailored value of Rn is 



1 



^^n]T -jT^ 



(9) 



(10) 



where 

Then equation (5) reduces to 



^ 



74 + 1 274 



1 



274 



74-1 



74-1 74-1 7i + 1 

74/2(74-0 



nil /mi 



n 1/2 



74(P//Pl) 



^ 



(1 +,// 1^2) 1/(74-1) 



(11) 



Equations ( 1 0) and (11) represent the tailoring conditions for thermodynamically perfect gases. 
From equation (11) it is evident that for a given combination of driver and driven gases, the ideal 
loading pressure ratio Pj/p^ required for tailored operation is a constant, regardless of the quantity 
of energy discharged into the driver gas. The loading pressure ratios for tailored operation with the 
different gas combinations used in these experiments are given in table 1 . 



The tailored shock Mach number is given by 

_ 2(^4 /a 1) 



[M 



s^T 



[Rn\ x 



(74-l)(l-e) 
Using equations (6) and (7) to obtain a^/ui , equation (12) can be written as 



(12) 



[Ms]j= [Rn^T 



(74-l)(l-e) 



74 ^1 
7i fn^ 



l+(74-l) 



vEc 
Vp 



-, \l/2 



I J 



(13) 



where e «i(7i - iy(7i + 1) is the perfect gas value of the density ratio across a normal shock wave 
for Ms » 1, and [Rn] f is the constant given by equation (10). The tailored shock Mach number 
[Ms] T depends on the initial loading pressure p^, the driver volume V, the stored capacitor energy 
Ec, and the choice of gases. A typical variation of [Ms] f with Pj , with He driving N2 , is given in 
figure 1, for £'c = 10^ J and V- 0.0209 m^ (Ames 2.90 m arc-heated driver). As shown, this result 
from equation (13) compares favorably with a similar result from reference 2 where, for the same 
He-N2 combination, nitrogen is treated as a real gas. 



'Ng real gas (ref. 2) 
17=1.0 



He-Ng 

Ec = 10^ joules 

V = 0209 m^ 



The preceding analysis indicates that if 
the appropriate ratio of p^ to p j is chosen the 
contact surface will be tailored at all shock 
Mach numbers. With the assumption of ideal 
shock-tube theory with thermodynamically 
perfect working gases inherent to this 
analysis, each tailored shock Mach number 
has a unique ratio of temperatures across the 
contact surface {T^IT2). In practice, this is 
not the case, however, because the 
temperature T^ is influenced by real-gas 
effects, viscous losses, heat transfer to the 
wall, and nonuniform distribution of energy 
within the driver chamber. The temperature 
T2 is affected by similar viscous and heating 
losses, real-gas effects, diaphragm opening, 
etc. Thus, these nonideal effects make it 

impossible to predict from ideal theory the ratio of pressures that will give a tailored condition. 

However, a near-tailored reservoir can be obtained by adjusting p^- and Pj , as discussed later. 




Figure 1.— Variation of tailored-shock Mach number with 
shock-tube loading pressure. 



Test Section Flow 

This section considers methods of evaluating flow properties in the high-velocity test stream 
issuing from a steady reservoir of quiescent, shock-heated gas. It is assumed v|:hat the reservoir gas is 
in equihbrium, with respect to both vibrational and chemical energies. In reference 13 a method is 
described for determining the Reynolds number, viscosity, and atom concentration in a 
hypervelocity nozzle from the measurement of stagnation-point heat-transfer rates to three probes 
placed side by side (the three-probe method). The physical nose dimensions of the probes should be 
chosen so that one probe measures the heating rate q^^ in the presence of an equilibrium (chemical) 
boundary layer, the second qfy. in a frozen shock layer, and the third qr^^ in the free-molecule flow 
regime. With a pitot-pressure measurement, in addition to these three heating rates, the following 
relationships can be used. (The derivation of these equations and the probe design considerations 
are given in reference 13. The present results are subject, of course, to the assumptions of the 
reference analysis.) The shock-layer Reynolds number is 






2m^ 



K^fr 



(14) 



where 



C = 0.64(7+ \y'^Pr 



l/2p^-O.6^-0.2 5 



The ratio of dissociation to total enthalpy is 






^eq'^eq ' 






■ .1/2 

"eq eq 



+ (Le 



0.52 _ 



)qf, 



r^fr^ 



(15) 



In equation (15) the two heat-transfer probes are assumed to be geometrically similar. Using the 
approximations 



one obtains 



Pt '^ PooUo, 



f/o, 



and -^oo *** 2 ^°°^ "*" ^/?"°° 




(16) 






Mp = 



2Al 



\dfr Pt qf r 
4 C^~ q}^' 



(17) 



(18) 



Hr. 



Ov 



2Le 



0. 52 



• .1/2 

^eq"eq 



kfrdf, 



II 



+ {Le 



0.52 



I) 



(19) 



U ^ 



_ ^eq eg _ . ' 



(20) 



All the flow properties in the preceding equations are expressed in terms of the measured quantities, 
stagnation heating rate and pilot pressure. The unknowns are the two transport properties, Lewis 
and Prandtl numbers, and the shock density ratio e, which must come from another source. 
Fortunately, they are relatively insensitive to test conditions and can be matched by an iterative 
solution if desired. An examination of equations (14) through (20) shows that the derived values of 
free-stream velocity U^, density p^, shock-layer viscosity /ig, and Reynolds number [Re] ^ require 
only two of the three measured heating rates, namely q^y and ^^^; and that the free-stream Mach 
number M^o and Reynolds number Re^ cannot be obtained by this method, since the static 
temperature of the stream is not directly available. However, these latter properties can be derived 
from the present measurements by an adaptation of the sudden-freeze method of reference 6, as 
vnh be shown later. To assess the validity of this combined approach, a comparison will be made 
between the stream densities from the three-probe and the sudden-freeze methods, while U^ from 
equation (16) will be compared to the stream velocity obtained by a completely independent 
analysis. 



EXPERIMENTS 



In this section the methods discussed for determining driver efficiency, tailored-interface 
conditions, and flow properties are apphed to the EAST facility. 



The Test Facility 

An electric-arc driver and energy storage system has been developed for shock-tunnel 
application, with provision for driver lengths up to 3 m. Such driver lengths are required to match 
the reservoir time requirements for tailored-interface operation (ref. 2). The energy for the arc is 
supplied by a 1250-/xF capacitor bank capable of storing 1 million joules when charged to a 
maximum of 40 kV. The arc is contained within a 10.2-cm-diam insulated chamber and is initiated 
by a tungsten wire. Details of the operating characteristics for the driver system are given in 
reference 14. (Dannenberg and Silva have developed methods to improve driver performance and 
range of operation; these are described in a forthcoming publication.) For the present experiments, 
driver lengths of 1.37 and 2.90 m were used. The energy per unit volume Ec/V was kept the same 
for both lengths (about 42 J/cm^) with capacitor bank voltages of 28 and 40 kV, respectively. The 
driver is coupled to a 10.2-cm-diam, 1 1.3-m-long shock tube, which discharges into a 10° half angle, 
conical supersonic nozzle. A 76-cm, hexagonal test section and a vacuum tank are located 
downstream of the nozzle. Figure 2 is a schematic diagram of the facility. Interchangeable sonic 




Driven tube 
I 1.3 m 



=^ 



20° _ Test ^ 
nozzle section 


Vocuum chamber 
4.3m 


0.76ni \ 
exit diQ/ 







rf/"^"^ 






J,___jj 



Collector 
ring 



Mom 
diaphragm 



Loading 
valve 



Nozzle 
diaphragm 



Model support 
spool 



Figure 2.— Schematic of Ames electric-arc shock tunnel. 

throats can be inserted at the nozzle entrance to permit operation at different flow Mach numbers. 
The shock wave arrival was monitored at six stations in the shock tube using the ion-probe system 
described in reference 15. Shock Mach number was calculated from the measured velocity and the 
initial gas temperature in the tube. The initial pressure p, was measured with a precision dial gage at 
the gas loading station, and the reservoir pressure po with a piezoelectric transducer located in a 
sidewall port at 0.64 cm from the endwall. Test gas impurities were estimated to be less than 
1.0X10"^ of the total sample. 



Driver Performance 



Driver efficiencies, determined by the method described earlier, are shown in figure 3 for the 
range of loading pressures p^ of the present tests. Equivalent driver temperatures computed from 
equation (6) are included for reference. Experimentally determined efficiencies were based on the 
shock velocity measured near the end of the 1 1.3-m tube and thus include the energy loss associated 
with shock attenuation. At values of p^ from 10 to 15 atm, the He and N2 drivers operate with an 
overall efficiency of about 50 percent. With the addition of argon to the driver gas, which lowers Mg 
because of the decrease in a^, the driver efficiency varied substantially from run to run. In all cases 
17 was better than for helium alone, but the wide variation from 0.60 to over 1.0 is a surprising 
result. It is inferred that the arc-discharge and heat-transfer processes in the driver chamber are 
variable, perhaps because of local variations in gas mixture ratio. Since the basis of comparison is a 



1.2 



1.0 



T4,°K 7500 6000 4500 



3000 




8 



Flogged symbols = 2.90 m driver 

1 ] 

24 32 



16 

Pi I atm 

Figure 3.— Overall driver efficiency and equivalent temperature; 
1.37- and 2.90-m drivers. 



uniformly heated driver gas, it is obvious 
that nonideal effects can be strong 
enough to create shock waves that 
exceed the ideal velocity. It should be 
noted, moreover, that rj > 1.0 has been 
reported in reference 8 for pure helium 
drivers at relatively low temperatures T^ . 
Here again one logical explanation is 
nonuniform driver heating. Further 
investigation of this behavior is 
warranted since it will influence 
interface temperatures T2 and T3 , which 
determine tailored operation. 



Reservoir Conditions 

In an earlier section, the ratios of initial driver-to-driven-tube pressures were predicted for 
tailored interface operation with ideal gases. For a given combination of gases the ratio is a fixed 
constant, although absolute values can be changed to vary Ms- As part of the experiment, an effort 
was made to validate this prediction, using the pressure ratios listed in table 1 . It was quickly found 
that nonideal effects resulted in undertailored conditions — an undertailored condition is a 
mismatch of internal energy at the contact surface such that E^ > £'2 , causing a downstream 
expansion wave to emerge from the reflected shock-contact surface interaction. This situation can 
be corrected by lowering pj, while keeping a^/ai constant, or by decreasing a^/ai with Pj 
constant, until the tailored condition is achieved. To obtain reasonably constant reservoir pressure 
histories with nitrogen as the test gas, it was necessary to increase the ratios p^/p ^ by about an order 
of magnitude by a reduction in p j . This is not surprising since the shock Mach numbers attained in 
most high-energy shock tubes are below ideal predictions at the initial pressures considered here. 
The observed departure from the ideal pressure ratio in this experiment is attributed to an elevated 
temperature behind the contact surface T^ , partly due to deceleration of the incident shock wave 
and partly to a timewise variation of total temperature in the gas issuing from the driver. These 
nonideal effects are the subject of a separate investigation. 

Figure 4 shows representative reservoir pressure histories for the Ms range and the driver gases 
used. Figure 4(a) is a record of a somewhat undertailored condition at Ms = 18.5, which has a fairly 
constant pressure after the arrival of the expansion wave from the contact surface. This pressure 
level corresponds to an enthalpy in the test gas of about 40 kJ/g, considerably above that obtainable 
in combustion or heated hydrogen driven facilities. The tailored condition would occur at a slightly 
lower Pj foT Ms> 1 8.5. Figures 4(b) and 4(c) show slightly overtailored conditions for which weak 
compression waves augment the reflected-shock pressure. 

An independent check of the reservoir gas quaUty was obtained by monitoring the radiant 
emission of the shock-heated nitrogen, using a photodiode detector with S-1 response. The reservoir 
gas was observed through a sidewall window masked by a 0.025-cm vertical slit for coUimation; 
signals were recorded on an oscilloscope. Although the detector was not calibrated for absolute 
intensity, care was taken to operate in the linear response range. Typical results (fig. 5) show that 



8 



U— JsOO/isec ^Expansion Wave 



li— ■■■ 

■■■hBR 




23.8 atm 
(a)Ms=i8.5 Pi= 10.2 atm (He), P|=:. 0092 atm (Ng) 

iWW Pim jPi 36.1 atm 

(■■■■T 






{b)Ms=l2.7 Pi=l2.2atm (.4A+.6He), P| =.OI32atm (Ng) 




25.6 atm 



{c)Ms=6.9, P|= 10.2 atm (Ns), P| = . 0264 atm (N2) J 

Figure 4.— Reservoir pressure histories at near-tailored 
conditions. 




(b)Ms = l5.4, P|=0.0I32 atm 

Figure 5 .— Pressure and radiation histories in the 
reservoir, 0.64 cm from end wall. 



the radiation from the reservoir varies with the pressure for at least 1 msec at Af^ = 6.9, and for 
about 0.5 msec at Ms = 15.4. In a quahtative sense, this finding indicates that no large dilution of 
the test gas occurs in these times. 



Test Section Measurements 



As the shock-heated, partly dissociated nitrogen in the reservoir expands into and beyond the 
throat of the supersonic nozzle, a point is reached where the recombination and vibrational 
relaxation become rate limited and can no longer follow the rapid change of temperature. When this 
occurs, the flow changes over a relatively short distance from local equilibrium to a state in which 
one or both of the two forms of internal energy (vibration and dissociation) are nearly invariant. In 
effect, the flow "freezes" at an intermediate level of dissociation and vibrational energy 
distribution, and remains essentially in this state as it sweeps through the test section. 
Consequently, the properties of the test stream differ from those in an equilibrium expansion; the 
temperature, pressure, and velocity are lower while the Mach number is higher. The calibration 
method must determine these differences, measure the dissociation level, and if possible, the 
vibrational energy defect of the flow. 

The three-probe rake— The flow quantities in the test section were evaluated, in part, 
using the three-probe method described earlier. The probes were mounted together on a single rake 
for simultaneous measurements of stagnation-point heating rates in the equilibrium boundary layer, 
frozen shock layer, and free-molecule flow regimes. The nose diameters were <ig„ = 5.1cm, 
dff. = 1.0 cm, and df^ = 0.10 cm; the largest probe was hemispherical and the two smaller probes 
were cylindrical in shape. Figure 6 shows a diagram and a photograph of the three-probe rake 
immersed in the test flow. For the present range of conditions, the computations presented in 
reference 13 show that a cylindrical probe of 0.10 cm diam is sufficiently small to obtain 



-0.5 cm radius cylinder 




(a) Rake assembly 




(b) Rake in fest stream 



Figure 6.— Three-probe heat-transfer rake. 



free-molecule flow at the stagnation point. On 
the basis of estimates and experimental results 
presented in reference 16, a cylindrical probe of 
1.0-cm diam should result in frozen shock-layer 
conditions. 

Estimates of the probe size for equilibrium 
boundary-layer flow in the stagnation region 
were based on the theory of reference 17. 
Although the reference analysis is specifically 
for air, it was assumed that the functional 
dependence of gas-phase reactions on local 
conditions would be similar for pure nitrogen. 
The resulting estimates of probe size, in the limit 
of a fully equilibrium boundary layer, exceeded 
the physical dimensions of the facility for many 
of the expected test conditions. Therefore, it 
was decided to relax the requirement of full 
equilibrium, retain the concept of three 
simultaneous measurements, and resolve the 
defect in heating rate by analytical means. A 
5.1-cm-diam hemisphere was used for this 
measurement; for convenience, this probe will 
be referred to as the equilibrium heat-transfer 
probe. 



(A:^P^c^)1/2 =0.153 J cm""2 °K~^sec" 



The sensing elements on the three probes were thin film platinum gages. To avoid shorting due 
to shock-layer ionization, all the gages were coated with a 1-^-thick layer of silicon monoxide by a 
vacuum evaporation technique. (The response time of a l-/i glass layer is about 50 jUsec for 
5 percent accuracy.) The backing material used was Pyrex 7740, which has the following properties: 

'^''^ and ap=2.18X10~3 fi/fl °K. The heat-transfer rates 
were determined directly with the use of analog 
networks. The methods of reference 16 were 
followed closely in the design and construction 
of both the gages and the analog networks. 

The survey rake— During test core sur- 
veys, other heating rates were measured on 
hemispherical models 2.5 cm in diam mounted 
on a rake spanning the test region (fig. 7). These 
models were copper shells, 0.013 cm thick, 
instrumented with 40 gage chromel-constantan 
thermocouples formed by soldering the 
individual wires of each pair into two closely 
spaced holes at the stagnation point. AU the 
heat-transfer data were recorded by a high-speed 
data acquisition system (ref. 18), which provides 
digital information at the rate of one reading 




13 cm 
adjustment , 



(a) Rake assembly 



(b) Rake in test stream 



Figure 7.— Pressure and heat-transfer survey rake. 



10 




Tg = Flow Starting time 
Test time 



2 3 4 

Time, ms 

Figure 8.— Rate of heat transfer to the equilibrium 
boundary -layer probe (Ms = 12.4, Pj = 0.0132 atm). 



every 2jLtsec. The data were punched on a paper 
tape and processed on an IBM 7094 computer. 
A sample of the machine-plotted heat-transfer 
data is shown in figure 8. 

Pitot pressures were measured, in repeat 
runs at each test condition, with a piezoelectric 
pressure transducer (Kistler, model 701 A). 
These measurements were obtained with both a 
flush mounting (gage face exposed to the flow) 
and a recessed mounting arrangement. In both 
cases the gage face was coated with a very thin 
layer of silicone rubber to reduce the heating 
effects. (Reservoir pressures were measured with 
similar transducers (Kistler, model 60 IH).) Pitot 
pressures were also measured during test core 
surveys with an adjustable rake, which contained 
as many as seven variable-capacitance pressure 
transducers. These gages were mounted behind a 



heat-sink baffle to shield the active element (a pretensioned Invar diaphragm, 0.0013 cm thick) 
from heating effects. 

The heat-transfer rates and the corresponding pitot pressures were measured for a wide range 
of shock Mach numbers (hence reservoir enthalpies) with nitrogen as the test gas. The driver gases 
were nitrogen, pure helium, and helium diluted with different amounts of argon; for each case the 
ratio of initial driver pressure to driven tube pressure Pj/p^ was adjusted until a near-tailored 
condition was achieved. The three-probe rake was mounted in the horizontal center plane at the 
exit of a conical supersonic nozzle with a 0.32-cm-diam throat. The pressure and heat-transfer 
survey rake was mounted vertically in the same nozzle, at the same axial station, and offset to 
position pitot probes at the same radial distance as the outer models on the three-probe rake. 
Throat sizes of 0.32 and 1 .27 cm were used with this rake. For the present range of experiments, 
the corrections to measured pitot pressures for rarefaction effects were estimated by the method of 
reference 19 and found to be less than 2 percent. The flow starting time in the nozzle (as in ref. 20, 
the time required to estabHsh uniform flow after arrival of initial disturbance) is estimated from 
probe response to be between 1 50 to 300 Msec, and in general, varies inversely with A/5. 



DISCUSSION 



The data are summarized in table 2 ; the measured heating rates, pressures, and test times are 
grouped into two sets according to the instruments used. Test 1 refers to the runs made with the 
three-probe heat-transfer rake, using thin-film detectors. Test 2 results are from the survey rake that 
used thin-wall calorimeters to measure heat-transfer rates and variable capacitance transducers for 
pressure. 

Analysis of the heat-transfer data from three-probe rake (test 1 ) gave inconsistent results over 
the range of test conditions. To isolate the cause of this erratic behavior, test 2 repeated several runs 

11 



at identical starting conditions and extended the scope to include a larger nozzle throat size, 
d* = 1.27 cm. Furthermore, as a check on the measuring techniques of test 1, a different type of 
heat-transfer sensor and pressure transducer were used. These were mounted in the survey rake that 
spanned the test region. The results of test 2 showed that the inviscid test core was smaller than 
predicted at the higher reservoir enthalpies. This result can be seen in 'table 2 where, for test 1, 
columns 7 and 1 indicate when the pitot pressure at the off-centerline location of the frozen-flow 
and free-molecule probes was lower than centerline pitot pressure. Column 13 of table 2 lists the 
time interval of useful data for each run, while column 14 gives an estimate of the duration of test 
gas flow based on reservoir radiation and test section heating rate measurements. In general, the 
longer flow times of column 1 4 could not be fully utilized because of pressure disturbances in the 
reservoir. 



Nozzle Test Core 

Rake surveys of the nozzle flow at the exit station, using the 0.32-cm-diam throat, showed 
that inviscid test cores varied in diameter from about 23 to 10 cm as the total enthalpy was 
increased from 8 to 40 kJ/g. Pressures were relatively constant across the core area and repeatable 
from run to run. At H^o > 27 kJ/g, the stream core size decreased to the point where the outer two 
probes on the heat-transfer rake (located 8.4 cm from the centerline) were in the outer edge of the 
nozzle wall boundary layer. For these conditions, the frozen-flow probe data were discarded, but the 
free-molecule probe results were assumed to represent a velocity close to the inviscid core value, 
since in the outer edge of a hypersonic boundary layer the density decreases much more rapidly 
than the velocity. For example, for the first run listed in table 2 the local pitot pressure was about 
0.7 the centerline value; the corresponding defect in velocity is estimated to be less than 1 percent, 
while the local density decrease is nearly 30 percent. The corresponding local Mach number is only 
about 1 5 percent lower, so that the free-molecule probe should remain within the appropriate flow 
regime. 

With a nozzle throat diameter of 1.27 cm, the measured test core diameter was greater than 
40 cm over the entire range of reservoir conditions. 



Total Enthalpy 

As shown in table 2, the stagnation region boundary layer on the largest probe, as expected, 
did not come to chemical equilibrium. Therefore, it was not possible to compute the stream total 
enthalpy directly using equation (19). An alternate method was devised for indirect verification of 
the enthalpy, using the measured q^^^ and q^^, along with an H^j derived from reflected shock 
conditions. The first step in this approach was to determine the chemical state of the shock layer by 
comparing the product of the rate of nitrogen dissociation and the flow residence time to the 
limiting value of equilibrium dissociation. The dissociation rate was estimated using the 
characteristic equation, as given for example in reference 21, which describes a diatomic gas with a 
binary collision mechanism 

doc AP -a IT .« , . 

— =(l-a) —— e-G'^ (21) 

dt Rp+CO 



12 



where 00= 1.5 is half the number of degrees of freedom contributing energy during a collision and 
A is a constant determined by experiment for a particular gas. In the present case da/dt was 
maximized by using the ideal gas (a= 0) temperature just behind the shock wave. 

Residence time was estimated from shock detachment distance and average local velocity. The 
uniform result was that even for this maximum dissociation rate the residence time was so short 
that the inviscid shock layer was effectively frozen at the stream composition (on the largest model) 
for all test conditions. Similarly, for the stagnation region boundary layer, a recombination rate 
parameter (the ratio of atom diffusion time across the boundary layer to characteristic atom 
lifetime) 

C, = L6X10iVd_ (22) 

2r*zi^2 

was used after the manner of reference 22, with pressure in atomspheres and nose diameter in 
centimeters. Values of Cj varied from 1 0~* to 1 0~^ over the test range and indicate atom lifetimes 
much in excess of transit time. Thus, the boundary layer in the stagnation region of the 5.1 -cm 
probe is frozen to atom recombination for all test conditions. 

With the chemical composition of the probe shock layer thus defined (frozen at stream 
composition) and assuming that reservoir enthalpy can be evaluated from the measured reservoir 
pressure at the entropy level associated with the reflected shock wave, the second step is to 
compute the convective heat transfer to a noncatalytic (glass covered) surface for several assumed 
values of a^. When the expression for a frozen free stream and frozen shock layer is used (ref. 22), 

where the heating rate to a fully catalytic surface q^^ is essentially the same as that for an 
equilibrium boundary layer q^ . As suggested in reference 22, this latter value can be closely 
approximated for nitrogen by 

^^'i \dj V 10,000/ V ^00/ 

where Uoo — (2^oo)^^^ . the flow velocity for an equilibrium expansion; that is, Uoo includes 
the dissociation energy. The resulting values of q ^ were interpolated to match the measured 
heating rate on the 5.1 -cm probe, which yields Ooo and the corresponding stream velocity Uoo- 

Alternately, U^o can be computed from equation (16) using the measured data for each test 
condition 

29/m 
U^= -^ (16) 

Pt 



13 



200 

i" 

^ 160 

I 

I 120 

C 

a> 

•o 80 

m 
o 

E 40 



Test 



I 
O 



Compufed from measured 
reservoir pressure at reflected 
shock entropy 




ref. 23 



8 12 16 

Stiock fvlach number, Mg 



Figure 9.- Variation of total enthalpy with Al^ for 
nitrogen. 



20 



The results of this analysis are shown in table 3 , 
where the two sets of Uoo are seen to be in good 
agreement. This finding implies that there are no 
significant energy losses in the reservoir for the 
range of conditions tested, and that the total 
enthalpy of the test core matches that of the 
reservoir. 



The variation of total enthalpy with Ms is 
shown in figure 9. Symbols denote computed 
values based on the measured reservoir pressure 
and the entropy level associated with the 
reflected shock wave. Computed enthalpies are 
compared with a theoretical, reflected shock 
curve taken from reference 23 ; differences are 
due to the change of reservoir pressure after 
shock reflection. 



Free-Stream Properties 

A nonequilibrium nozzle flow program similar to that of reference 9 was used for comparing 
the measured flow properties with theoretical predictions. The major uncertainty involved in this 
program lies in the assumption that the vibrational energy mode always remains in equilibrium with 
translational and rotational modes. (While the energy in vibrations is at most about 20 percent of 
the total enthalpy, its disposition can make a noticeable difference in stream properties, as will be 
shown later.) The reservoir temperature and pressure are specified as initial values for the numerical 
solution. In the present cases the initial values were the measured reservoir pressure and the 
corresponding equilibrium temperature obtained from thermodynamic charts for nitrogen, using the 
appropriate reservoir enthalpy. Energy losses from the reservoir gas were assumed to be negligible; 
thus, the computed quantities represent a limiting set for the entropy associated with each test 
condition. The numerical computations showed that the nitrogen atom concentration was frozen in 
the test section for all the cases considered. 

The semiempirical method of reference 6 also can be used to predict certain stream properties, 
without knowledge of the specific relaxation processes or rate constants. It is assumed that the 
gasdynamic properties at the test station (having undergone relaxation of several internal degrees of 
freedom) are approximately the same as if the gas had made a sudden transition from equihbrium to 
a zero rate of internal energy exchange at some location in the expansion. The location of the 
transition is characterized by a single parameter, the freeze Mach number. Downstream of this point 
both the chemical composition and vibrational energy distribution of the gas are constant (frozen). 

A computing program for one-dimensional nozzle flow, using equilibrium reservoir conditions 
as input quantities, was used to generate stream properties as a function of area ratio for a specified 
set of freeze Mach numbers. The pitot pressure was calculated assuming the vibrational mode and 
chemistry remained frozen behind the bow shock wave — a realistic choice for the flow conditions 
of the present experiment. The details and an experimental verification of this method are given in 
reference 6. 



14 



Vibrational 






energy 

Frozen Equil. 

• O 

■ D 


d« 


cm 
0.32 
1.27 



Kest 2 




38 42 

Reservoir entropy, Sq/R 



Figure 10.— Free-stream atom concentration. 



Atom concentration— The measured and 
computed values of (x^ are compared in 
figure 10 as a function of reservoir entropy. 
Reservoir atom concentration a^ is common to 
both theory and experiment. The open symbols 
represent concentrations deduced from 
measured heat transfer, using equations (23) and 
(24) as described earlier, with vibrational energy 
assumed to be in equilibrium. For test 2 the 
probe surface was partly catalytic (copper 
oxide), and allowance was made for surface 
recombination by the method of reference 24. 
The recombination energy fraction varied from 
4 to 26 percent of the available dissociation 
energy. 



The filled symbols in figure 10 represent a similar analysis in which both vibrational and 
chemical energy exchanges become inactive simultaneously in the nozzle expansion and remain so 
in the probe shock layer. Comparisons of shock-layer residence time with vibrational relaxation 
rates showed this a reasonable assumption except at the highest enthalpy conditions. Tables 4 and 5 
summarize reservoir conditions, stream atom concentrations, and other flow quantities. 

The results of figure 10 indicate that a relatively small amount of recombination occurs in the 
expanding nitrogen flow over most of the operating range, and that the numerical solution (with 
vibrational equilibrium) substantially underestimates the stream atom concentration. For frozen 
vibrations, the sudden-freeze approximation indicates that the internal energy adjustment stops just 
downstream of the nozzle throat, between M^= 1 and 2. The effect of nozzle throat size on atom 
concentration appears relatively minor over the test range. At SqIR > 42, recombination is 
enhanced and ol^ approaches the numerical prediction. 



86 



J) 5 

E 




34 



^ 


D d< i:27;Test 2 

O d* -- 0.32 Test 1 

Filled symbols denote 
frozen vibrations 


38 

Reservoir 


42 46 50 
entropy, Sq/R 



Figure 1.— Comparison of measured and predicted 
stream velocity. 



Velocity- Experimental and predicted 
free-stream velocities are compared in figure 1 1 . 
The sudden-freeze predictions are shown as 
dashed lines, shifted downward slightly relative 
to the numerical solution and the a^ curve to 
account for the frozen vibrational energy. The 
experimental values are consistent with the atom 
concentrations of figure 10 and, as expected, are 
substantially lower than the numerical 
prediction. At the lowest values of SoJR the 
data fall below the curve for a^ = a.^ because of 
frozen vibrational energy. At the higher values 
of Sq/R, the velocities move toward the 
numerical prediction. 

In the present analysis, the flow velocity 
was derived from experimental measurements in 
two ways: (1) using the stagnation heating rate 



15 



20 



cm 
O d* = 0.32\ ^ , „ 
D d*= 1.27; ^«=' 2 
O d*n 0.32 Test I 
Filled symbols-eqn. 17 
Open symbols -method of ref. 6 
Flagged symbols- frozen vibration 
Symbol superscripts = po, otm 



= 2500 



Computed -ref. 9 



for a frozen shock layer, and (2) using the 
free-molecule heating rate as in equation (16). 
Comparisons are given in table 3. The first 
method requires a prior knowledge of the stream 
total enthalpy; the second does not. For this 
reason, the ^f^ rn^^^od is preferable and, in 
fact, is a relatively simple way to determine the 
velocity of a nonequilibrium stream. For best 
accuracy, both q.c^ and p^ should be measured 
simultaneously. 

Density— The computed and measured 
stream densities are shown in figure 12, where 
the filled symbols represent values from 
equation (17) and the open symbols are from 
the sudden-freeze analysis. The curves are from 
the exact numerical solutions (with vibrational 
equilibrium) at two representative nozzle area 
Figure 12.- Free-stream density. ratios A I A* and reservoir pressures p^. The 

experimental results agree in general with predictions on this summary plot; more exact 
comparisons would require a point-by-point analysis since p^o is sensitive to both reservoir pressure 
and local area ratio. (The area ratios shown are average values for each d* as determined from the 
sudden-freeze analysis, while p^ is within 1 percent of experiment except for the symbols with 
superscript values. For more detail see tables 4 and 5). The results for test 1 , using the sudden-freeze 
analysis, compare favorably with the more direct values from equation (17). This agreement 
between stream densities appears to justify the assumption that the sudden-freeze analysis can be 
used to supplement the stream properties obtained by the three-probe method. 




Shock-layer viscosity— The viscosity in the stagnation-region shock layer is obtained from 
the relationship of pitot pressure and heating rates given in equation (18) 



^e ATI 



AO 



^fm 



(18) 



where 



C= 0.64(7 -1- 1)1/ 2^^-0.6^-0 .2 5 



The density ratio across the bow shock e was evaluated using either 7 = (8 -i- )/(8 - ) for 
equihbrium vibrations, or 7 = (4 -I- 3 )/(4 + ) for frozen, in the normal shock relation 
PooIPp ^ (7 ~ l)/(7 + 1) for Mao >^ 1 ■ A Prandtl number of 0.7 was used throughout. The results 
are listed in table 4, where the values are seen to increase with H^ in a regular manner. As a check 
on the magnitude of these values, equation (A4) of reference 24 was used to determine the 
shock-layer temperature T^. Figure 13 shows that the resulting temperatures exceed the estimated 
values for dissociated flow with equihbrium vibrational energy; thus the measured viscosities appear 
consistent with the other findings of this investigation. 



16 



20 



15 



10 



Test I 
Symbols are for T^ from 

72 + 1 



[^e = l 58X10"^ Te°^' 



28 dco 



I + aco 



ref 24, eqn (A4) 




30 



34 



38 42 

S„/R 



46 



50 



Shock-layer Reynolds number— Reynolds 
numbers in the stagnation - region shock 
layer were obtained from equation (14) and are 
listed in table 4. They remain fairly constant 
over the operating range, varying from 4.2 to 
6.9/cm, as a result of the relatively small changes 
in density and viscosity observed earlier. These 
low Reynolds numbers reflect the high viscosi- 
ties in the frozen shock layers of the present 
experiments. 

Much number, Reynolds number, and 
A/A*— The sudden-freeze method was used 
to evaluate the additional flow quantities, 
free-stream Mach number, Reynolds number, 
and effective area ratio. Results are listed in 
tables 4 and 5. For the smaller throat size 
(0.32 cm) the Moo increased with total enthalpy 
from 20 to 32 and for the 1.27-cm throat from 
14 to 23. The Reynolds number did not vary in 
a consistent manner with increasing enthalpy, apparently because of the increase in freeze Mach 
number (see fig. 1 1); values between 200 and 1000/cm were obtained. A substantial difference was 
observed, at intermediate enthalpies, between the Reynolds numbers for frozen and equilibrium 
vibrations (see table 5). Thus, a knowledge of the vibrational energy distribution is important in 
evaluating the properties of nonequilibrium streams at these energy levels. 

The effective area ratio of the test stream comes directly from the sudden-freeze analysis, as 
shown in reference 6. The ratio A/A* decreased with increasing reservoir enthalpy from 12,000 to 
about 9,000 with the small nozzle throat and from about 3,300 to 2,300 with the large throat. (The 
corresponding geometric area ratios are 56,000 and 3,600.) Representative boundary -layer 
displacement thicknesses 6* at the nozzle exit are listed in the last column of table 5. The very 
substantial 20- to 23-cm thicknesses that occur with the small nozzle throat correspond to the small 
inviscid core size noted earlier. Obviously, for the present reservoir pressure, 0.32 cm is about the 
smallest throat diameter that can be used in this nozzle. 



Figure 13.^ Shock-layer temperature. 



CONCLUDING REMARKS 



The operating characteristics of an electric-arc-heated driver and shock-tube system, and the 
properties of high energy nitrogen flows expanded in a supersonic nozzle have been investigated. 
Driver temperatures up to 8000° K and pressures up to 340 atm were used to obtain 
tailored-interface conditions at shock Mach numbers from 7 to 19, enthalpies in the reflected-shock 
reservoir of test gas from 7 to 40 kJ/g, and test-section Mach numbers from 14 to 32. The 
gasdynamic processes in the shock tube and nozzle have been compared with theoretical, real-gas 
predictions. Local equilibrium conditions were assumed to exist everywhere in the shock tube; the 
nozzle flow was not so restricted, and substantial nonequilibrium effects were observed. The 
primary results of this investigation are summarized as follows. 



17 



1. The overall efficiency of the driver and shock-tube system was found to vary with the 
composition of the driver gas. Argon-heUum mixtures were the most effective but did not give 
consistent shock-tube performance. Pure helium and pure nitrogen were less effective driver gases 
but produced mor^ uniform results. Efficiencies from 40 to 100 percent were recorded. 

2. An ideal, theoretical analysis of tailored-interface, shock-tube operation, which predicts a 
constant ratio of initial driver pressure to initial shock-tube pressure for tailoring at all shock 
velocities, was not confirmed by experiment. However, the nonideal effects that invalidated this 
prediction could be offset by increasing the pressure ratio, so that near-tailored operation was 
achieved for shock Mach numbers from 7 to 19. 

3. Reflected shock relationships and a reservoir pressure history can be used to define reservoir 
conditions at total enthalpies between 7 and 40 kJ/g, if successive pressure changes are 
considered to be isentropic. There was no significant loss of energy from the reservoir gas during 
the 0.5- to 1.5-msec test period. 

4. The proposed method of evaluating the properties of a nonequilibrium nozzle flow of nitrogen, 
using simultaneous measurements of stagnation heating rates (to noncatalytic surfaces) in three 
gasdynamic regimes, could not be fully utilized under current test-section conditions. Limitations 
on probe size prevented the attainment of an equilibrium boundary layer in the stagnation 
region. In future work, however, it may be possible to use a catalytic probe surface to recover the 
dissociation energy of the shock layer. 

5. By combining elements of the three-probe diagnostic method with a sudden-freeze 
approximation of the flow behavior it is possible to define most of the properties needed for 
hypervelocity, gasdynamic studies in a nonequilibrium stream. This technique has been shown to 
give consistent results in a single-component gas (nitrogen) over a wide range of specific energies, 
with initial dissociation up to 75 percent. 

6. Theoretical predictions were found to underestimate the energy retained in the inert degrees of 
freedom (dissociation and vibration) in a rapidly expanding stream of nitrogen. Typically, the 
flow departed from thermodynamic equilibrium just downstream of the nozzle throat and was 
effectively frozen at a local Mach number of about 2. 



Ames Research Center 

National Aeronautics and Space Administration 
Moffett Field, Calif., 94035, March 21, 1972 



REFERENCES 



1. Wittliff, C. E.; Wilson, M. R.; and Hertzberg, A.: The Tailored-Interface Hypersonic Shock 
Tunnel./. Aerospace Sci., vol. 26, 1959, pp. 219-228. 

2. Loubsky, W. J.; and Reller, J. O., Jr.: Analysis of Tailored-Interface Operation of Shock Tubes 
With Helium-Driven Planetary Gases. NASA TN D-3495, 1966. 

3. Hertzberg, A.; Wittliff, C. E.; and Hall, J. G.: Development of the Shock Tunnel and Its 
Application to Hypersonic Flight. ARS Progress in Astronautics and Rocketry; Hypersonic 
Flow Research, F. R. Riddell, ed., Academic Press, Inc., vol. 7, 1962, pp. 701-758. 

4. Nagamatsu, H. T.; and Martin, E. O.: Combustion Investigation in the Hypersonic Shock 
Tunnel Driver Section./. Appl. Phys., vol 30, July 1959, pp. 1018-1021. 

5. Mason, R. P.; and Reddy, N. M.: Combustion Studies in the UTIAS Hypersonic Shock Tunnel 
Driver. Proc. 5th Shock Tube Symposium, U. S. Naval Ordnance Laboratory, April 1965. 

6. Hiers, R. S., Jr.; and Reller, J. O., Jr.; Analysis of Nonequihbrium Air Streams in the Ames 
1-Foot Shock Tunnel. NASA TN D-4985, 1969. 

7. Camm, J. C; and Rose, P. M.: Electric Arc-Driven Shock Tube. Phys. of Fluids, May 1963, 
pp. 663-677. 

8. Warren, W. R.; Rogers, D. A.; and Harris, C. J.; The Development of an Electrically Heated 
Shock Driven Test Facility. Second Symposium on Hypervelocity Techniques, Univ. of 
Denver, March 1962. 

9. Lordi, J. A.; Mates, R. E.; and Moselle, J. R.: Computer Program for the Numerical Solution of 
Nonequilibrium Expansions of Reacting Gas Mixtures. NASA CR-742, 1966. 

10. Reddy, N. M.: Shock-Tube Flow Analysis With a Dimensionless Velocity Number. NASA TN 
D-5518, 1969. 

1 1. Glass, I. I.; and Hall, J. G.; Handbook of Supersonic Aerodynamics. Shock Tubes (Section 18). 
NAVORD Rep. 1488, vol. 6, Dec. 1959. 

12. Flagg, R. F.; Detailed Analysis of Shock Tube Tailored Conditions. RAD-TM-63-64, AVCO 
Corp., Wilmington, Mass., Sept. 1963. 

13. Reddy, N. M.; A Method for Measuring Reynolds Number, Viscosity, and Atom Concentration 
in Hypervelocity Nozzles. AIAA J., vol. 6, July 1968, pp. 1398-1400. 

14. Dannenberg, R. E.; and Silva, A. F.; Exploding Wire Initiation and Electrical Operation of a 
40 kV System for Arc-Heated Drivers up to 10 Feet Long. NASA TN D-5126, 1969. 



19 



15. Dannenberg, R. E.; and Humphry, D. E.: Microsecond Response System for Measuring Shock 
Arrival by Changes in Stream Electrical Impedance in a Shock Tube. Rev. Sci. Instru., vol. 39, 
Nov. 1968, pp. 1692-1696. 

16. Reddy, N. M.: The Use of Self-Calibrating Catalytic Probes to Measure Free-Stream Atom 
Concentrations in a Hypersonic Flow. NASA CR-780, 1967. 

17. Inger, G. R.: Nonequilibrium Stagnation Point Boundary Layers With Arbitrary Surface 
Catalycity..4/y4y4/., vol. 1, 1963, pp. 1776-1784. 

18. Seegmiller, H. L.; and Mazer, L.: A 500,000 Sample per Second Digital Recorder for the Ames 
Electric-Arc Shock Tunnel. IEEE Pub. 69 C 19-AES, 1969, pp. 243-247. 

19. Potter, J. L.; and Bailey, A. B.: Pressures in the Stagnation Regions of Blunt Bodies in Rarefied 
Flow. AIAA J., vol. 2, April 1964, pp. 743-745. 

20. Dunn, M. G.: AppHcation of Microwave and Optical Diagnostic Techniques in Shock-Tunnel 
Flows. AIAA Paper 68-394, 1968. 

21. Hammit, A.: The Flow of a Dissociating Gas Around and Behind a Blunt Hypersonic Body. 
BSD-TDR-62-107, May 1962. 

22. Pope, Ronald B.: Stagnation-Point Convective Heat Transfer in Frozen Boundary Layers. 
AIAA J., vol. 6, no. 4, April 1968, pp. 619-626. 

23. Lewis, C. H.; and Burgess, E. G., Ill: Charts of Normal Shock Wave Properties in Imperfect 
Nitrogen. AEDC-TDR-64-104, Arnold Engineering Development Center, Tenn., May 1964. 

24. Okuno, A. F.; and Park, Chul: Stagnation-Point Heat Transfer Rate in Nitrogen Plasma Flows: 
Theory and Experiment. ASME Pub. 69-WA/HT-49, Nov. 1969. 



20 



TABLE 1 .- RATIOS OF INITIAL DRIVER PRESSURE TO DRIVEN TUBE PRESSURE FOR TAILORED OPERATION 



Driver/driven gases 


He/Nj 


(0.3A + 0.7He)/N2 


(0.4A + 0.6He)/N2 


A/N2 


N2/N2 


PilPi 


117 


31.5 


25.5 


11.7 


25.0 



TABLE 2.- SHOCK TUBE, NOZZLE RESERVOIR, AND TEST STREAM QUANTITIES FOR NITROGEN 



1 


2 


3 


4 


5 


6 


1 


8 


9 


10 


11 


12 


- 13 


14 


Test 


P.XIO^, 


^s 


Po- 


^0' 


t?*, 


t 


Theoretical 


Measured 


Local 


Measured 


Measured 


Test 


Test flow 


atm 


atm 


kJ/g 


cm 


PfXlo^ 

atm 


'^eq 


"ieq 


P^XIO^ 


¥ 


'Ifm 


interval 


duration 














W/cm^ 


atm 


W/ 


cm^ 


msec 


1 


9.22 


18.5 


41.1 


40.6 


0.32 


6.94 


197 


127 


4.97 


116 


161 


0.5 










17.8 


43.5 


38.7 






8.03 


191 


119 


5.24 


127 




.6 








, 


17.7 


39.1 


38.5 






7.21 


181 


123 


4.62 


120 


153 


.5 


0.5 






13.2 


16.0 


40.8 


30.6 






6.12 


142 


— 


6.12 


119 


165 


.4 


.4 










15.3 


38.8 


28.8 






6.18 


134 


74 


5.37 


108 


131 


.4 


.4 










14.9 


38.1 


27.0 






5.71 


124 


68 


5.71 




136 


.7 












13.0 


35.4 


21.6 






5.30 


100 


58 


5.30 


91 


125 


.6 


1.2 










11.4 


39.5 


17.9 






5.24 


83 


55 


5.24 


83 


108 


.9 








' 


' 


11.3 


34.7 


17.1 






4.56 


75 


49 


4.56 


79 


108 


.7 


— 


f 


26.3 


6.9 


19.4 


6.9 






2.58 


22 


18 


2.58 


28 


40 


1.2 


1.6 


2 


13.2 


15.8 


59.8 


31.8 






9.80 


265 


162 








1.2 


1.5 






13.2 


13.7 


42.5 


24.4 






6.73 


177 


111 








1.2 


1.7 






26.3 


7.6 


22.1 


8.2 


T 


2.79 


39 


29 








.9 


1.3 






9.22 


17.9 


44.2 


39.2 


1.27 


22.6 


437 


341 








2.0 


2.4 






13.2 


16.2 


47.6 


31.7 






36.6 


475 


388 








1.2 


2.5 








12.3 


34.0 


19.3 






23.7 


259 


184 








1.4 


2.8 






■ ' 


11.6 


36.7 


17.9 






23.1 


236 


185 








1.4 


2.2 




' 


26.3 


7.3 


22.1 


7.8 


\ 




11.7 


77 


71 









2.0 


2.4 



to 



Measured 


Calculated 


Po' 
atm 


He 

kJ/g 


41.1 


40.6 


43.5 


38.7 


39.1 


38.5 


40.8 


30.6 


38.8 


28.8 


38.1 


27.0 


35.4 


21.6 


39.5 


17.9 


34.7 


17.1 


19.4 


6.9 



TABLE 3.- EVALUATION OF STREAM TOTAL ENTHALPY, TEST 1 



Equivalent 

0.57 

.58 

.53 

.50 

.47 

.43 

.30 

.22 

.20 






Measured 




% 


Qfr' 
W/cm^ 


E( 


0.74 


127 




.68 


119 




.69 


123 




.51 


119 




.48 


74 




.44 


68 




.32 


58 




.22 


55 




.20 


49 




.02 


18 





Equivalent 


Measured 


Equivalent 


m/sec 


W/cm^ 


(f4o)/;„' 

misec 
6430 


6590 


161 


6220 


— 


— 


6480 


153 


6450 


5260 


165 


5310 


5120 


131 


4800 


5030 


136 


4710 


4800 


125 


4650 


4590 


108 


4600 


4560 


108 


4660 


3290 


40 


3160 



(£4c)/, 



m 



0.977 

.996 

1.008 

.938 

.937 

.970 

1.003 

1.022 

.962 







TABLE 4.- F 


LOW( 


:ONDITION 


fS AND TE 

PcoXlO^, 


ST-SECTIC 


N FLOW I 

M^XIO*, 


'ROPERTIES, TEST 1 

Sudden freeze 






P^, 


<L 


H „ 
















PooX10^ 


^s 


^0' 

atm 


p^xlo^ 

atm 


kJ/g 


So/R 


% 


Ooo 


Kg/m^ 
1.78 


[Re/cm]^ 


Nsec/m^ 


A/A* 


M^ 


[Rel cm] ^ 


kg/m^ 
2.27 


18.5 


41.1 


6.94 


40.6 


45.1 


0.74 


0.57 


4.49 


2.61 


8,900 


29.8 


289 


17.8 


43.5 


8.03 


38.7 


44.1 


.68 


.58 


2.10 


4.98 


2.61 


8,600 


31.8 


328 


2.59 


17.7 


39.1 


7.21 


38.5 


44.4 


.69 


.53 


1.74 


4.17 


2.70 


9,000 


29.3 


233 


2.20 


16.0 


40.8 


6.12 


30.6 


41.5 


.51 


.50 


2.25 


5.54 


2.19 


9,700 


32.2 


446 


2.58 


15.3 


38.8 


6.18 


28.8 


40.8 


.48 


.47 


2.40 


5.68 


2.20 


9,500 


30.1 


436 


2.92 


14.9 


38.1 


5.71 


27.0 


40.2 


.44 


.43 


2.29 


5.32 


2.15 


10,000 


29.4 


370 


2.45 


13.0 


35.4 


5.30 


21.6 


38.2 


.32 


.30 


2.33 


5.71 


2.03 


10,300 


25.0 


318 


2.67 


11.4 


39.5 


5.24 


17.9 


36.5 


.22 


.22 


2.51 


6.30 


1.85 


11,800 


24.5 


318 


2.93 


11.3 


34.7 


4.56 


17.1 


36.3 


.20 


.20 


2.23 


6.30 


1.64 


11,800 


24.0 


266 


2.78 


6.9 


19.4 


2.58 


6.9 


31.8 


.02 





2.45 


6.86 


1.22 


11,800 


19.5 


207 


2.22 



22 



ta 












TABLE 5.- 


- FLOW CONDITIONS AND TEST-SECTION FLOW PROPERTIES, TEST 2 








'2. 


X atm cm 


P^xlo^ 

atm 


kJ/g "^o/^ 


% 


Theoretical 
g g ^ a Measured 


Theoretical 
F-B-L.*^ 
E.V.'^ 

W/cm' 


Theoretical 
F.B.L.*^ 
F.V.'^ 

%\ 
W/cm^ 


E.V.*^ 
F.V.'^ 




Sudden freeze approximation 


CO 
-3 

to 


A/A* 


[i?e/cm]^ 


Moo 


PooX10^ 

kg/m^ 


5*, cm 




15.8 59.8 


0. 
1. 


32 

r 

27 

1 


9.80 


31.8 41.3 


0.53 


309 162 


158 
173 


157 
171 


0.50 
.46 


8900 
9600 


832 
640 


31.5 
29.6 


4.19 
3.94 


22.9 
22.4 


4i> 


13.7 42.5 


6.74 


24.4 39.0 


.36 


192 101 


109 
115 


110 
116 


.33 

.27 


11,000 


303 


24.0 


2.78 


21.3 




7.6 22.1 


2.79 


8.2 32.4 


.02 


41.0 29.2 


37.6 


29.6 


.02 


12,700 


169 


19.8 


2.10 


20.1 




17.9 


44.2 


22.5 


39.3 44.4 


.70 


574 341 


330 
346 


... 


.56 


3200 


705 


22.7 


6.75 


2.29 




16.2 


47.6 


36.6 


31.8 


41.7 


.53 


590 388 


377 
407 


... 


.40 


2250 


452 


17.5 


12.4 


8.14 




12.3 


34.0 


\ 


23.7 


19.3 


37.4 


.26 


287 


184 


178 
193 


179 
192 


.25 
.19 


2400 


675 


15.7 


12.0 


7.12 




11.6 


36.7 


23.1 


17.9 


36.6 


.22 


261 


185 


178 
186 


179 
186 


.19 

.13 


2600 
2800 


981 
518 


17.0 
15.2 


13.6 
10.4 


5.84 
4.57 




7.3 


22.1 


11.7 


7.8 


32.2 


.02 


80.6 


71.1 


80.6 


71.0 


.02 



3300 


453 


13.7 


9.48 


1.78 



^E.B.L. = equilibrium boundary layer 
F.B.L. = frozen boundary layer 

''E.V. = equilibrium vibrations 
F.V. = frozen vibrations 



to 



NATIONAL AERONAUTICS AND SPACE ADMISTRATION 
WASHINGTON, D.C. 20546 



OFFICIAL BUSINESS 

PENALTY FOR PRIVATE USE 1300 



FIRST CLASS MAIL 



POSTAGE AND FEES PAID 

NATIONAL AERONAUTICS AND 

SPACE ADMINISTRATION 




NASA 451 



021 001 CI U 11 720609 SC0gG3DS 

DEPT OF THE AIR FORCE 

AF WEAPONS LAB (AFSC) 

TECH LIBRARY/WLOL/ 

ATTN: E LOU B0Hr-5AN, CHIEF 

KIRTLAND AFR N"1 37117 



POSTMASTER: 



If Uttdelivewble (Section 158 
Postal Manual ) DoNotRetutn 



"The aeronautical and space activities of the United States shall be 
conducted so as to contribute . . . to the expansion of hu?iian knowl- 
edge of phenomena in the atmosphere and space. The Adjninistration 
shall provide for the widest practicable and appropriate dissemination 
of information concerning its activities and the results thereof," 

— National Aeronautics and Space Act of 1958 



NASA SCIENTIFIC AND TECHNICAL PUBLICATIONS 



TECHNICAL REPORTS: Scientific and 
technical information considered important, 
complete, and a lasting contribution to existing 
knowledge. 

TECHNICAL NOTES: Information less broad 
in scope but nevertheless of importance as a 
contribution to existing knowledge. 

TECHNICAL MEMORANDUMS: 
Information receiving limited distribution 
because of preliminary data, security classifica- 
tion, or other reasons. 

CONTRACTOR REPORTS: Scientific and 
technical information generated under a NASA 
contract or grant and considered an important 
contribution to existing knowledge. 



TECHNICAL TRANSLATIONS: Information 
published in a foreign language considered 
to merit NASA distribur .1 in English. 

SPECIAL PUBLICATIONS: Information 
derived from or of value to NASA activities. 
Publications include conference proceedings, 
monographs, data compilations, handbooks, 
sourcebooks, and special bibliographies. 

TECHNOLOGY UTILIZATION 
PUBLICATIONS: Information on technology 
used by NASA that may be of particular 
interest in commercial and other non-aerospace 
applications. Publications include Tech Briefs^ 
Technology Utilization Reports and 
Technology Surveys. 



Details on the availability of these publications may be obtained from: 
SCIENTIFIC AND TECHNICAL INFORMATION OFFICE 

NATIONAL AERONAUTICS AND SPACE ADMINISTRATION 

Washington, D.C. 20546