You are on page 1of 28

AIAA Propulsion and Energy Forum 10.2514/6.

2020-3812
August 24-28, 2020, VIRTUAL EVENT
AIAA Propulsion and Energy 2020 Forum

Thrust Measurement of a Hydrogen Peroxide Vapor Propulsion


System

Brandie L. Rhodes ∗
The Aerospace Corporation, El Segundo, CA, 90245

Evan R. Ulrich †
The Aerospace Corporation, El Segundo, CA, 90245

Andrea G. Hsu ‡
The Aerospace Corporation, El Segundo, CA, 90245
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Paul D. Ronney §
University of Southern California, Los Angeles, California, 90089

Direct thrust measurements were completed on a novel hydrogen peroxide (H2 O2 ) vapor

thruster utilizing a torsional thrust stand. Liquid H2 O2 temperature, nozzle exit angle, and

nozzle contact design were varied to determine the impact to performance. Specific impulse

was calculated using the measured thrust and mass flow rate. Results were compared with

an equivalent water vapor system and demonstrated a 60% improvement in Isp . Both H2 O2

and water thrusters showed losses associated with boundary layer formation within the nozzle.

Additional losses due to thermal contact with the manifold were witnessed in the H2 O2 version.

Modifications to the nozzle design and isolation are expected to lead to further performance

improvements.

I. Introduction

Liquid hydrogen peroxide (H2 O2 ) has been used as a monopropellant in space applications for decades, including use in

the V2 rocket turbopump as well as in the reaction control systems on the Mercury space capsule and the Soyuz descent

module [1]. Interest in H2 O2 faded upon the introduction of hydrazine in the late 1960s, a monopropellant with higher

specific impulse and higher toxicity. However, recently “green” propellants like H2 O2 have seen a resurgence, with

satellite manufacturers looking for less toxic and more environmentally-friendly solutions. Green propellants are also a

better fit with cubesat and small satellite systems, where (1) manufacturers might not have the infrastructure to handle
∗ Engineering Specialist, iLab, P.O. Box 92957, M1/195, AIAA Member
† Engineering Manager, Embedded Control Systems, P.O. Box 92957, M4/906
‡ Sr. Scientist, Propulsion Sciences, P.O. Box 92957, M2/341
§ Professor, Dept. of Aerospace and Mechanical Engineering, Viterbi School of Engineering, 3650 McClintock Ave. OHE 430J, AIAA Associate

Fellow

$copyRight
highly-toxic propellants and (2) additional safety requirements are imposed by rideshare scenarios.

The propulsion system tested and evaluated in this paper utilized vacuum-evaporated H2 O2 vapor as a monopropellant.

Like its liquid counterpart, H2 O2 vapor will also decompose on a catalyst and generate water vapor and oxygen according

to the following equation:

1
𝐻 2𝑂 2 → 𝐻 2𝑂 + 𝑂 2 (1)
2

Due to the low vapor pressure of H2 O2 (< 60 Torr for temperatures < 80 °C), the propellant flow rate and therefore thrust
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

associated with a vacuum-evaporated system is much lower than a traditional liquid system - targeting the millinewton

thrust range needed for attitude control of small satellites.

A. Background

The first prototype (shown in Figure 1) and the supporting theory were presented in Reference 2 - proving that a low

pressure, vacuum-evaporated H2 O2 vapor would decompose on a catalyst and produce high temperature gas useful for

propulsion. It also provided information on the required vapor mole fraction and catalyst temperature to initiate rapid

reaction. The second prototype, shown in Figure 2 and detailed in Reference 3, focused on material compatibility and

catalyst selection. Both were evaluated on catalyst temperature and pressures alone. Prototype 3, the version presented

in this study, made further modifications to increase performance and measured thrust directly.

Fig. 1 Prototype 1: test unit (left), CAD model (right).

2
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 2 Prototype 2: test unit (left), CAD model (right).

B. Design and Construction

Prototype 3 can be seen in Figure 3. One all aluminum GEMS solenoid valve with Viton® seals and a 1.32 mm orifice

was used to control vapor flow. TE Connectivity pressure sensors were used to measure tank pressure upstream of the

valve. One for ambient conditions (range of 0 - 100 psi) and the other for higher resolution at operating conditions

(range of 0 - 5 psi). The error band of these sensors is ±3% span and they use a SS 316L diaphragm as the sensing

element with a Viton® o-ring seal. For one variant of the design, the pressure was also taken downstream of the catalyst

ahead of the nozzle. This was done with a Honeywell TruStabilityTM pressure sensor with a 0 - 1 psi range (error 0.25%

span). Due to its location, downstream of the catalyst, the materials did not have to be highly compatible with H2 O2 .

Therefore a sensor was chosen with a tighter range and smaller size. Heaters and temperature sensors were placed on

the outside of the tank and manifold to maintain desired liquid temperatures during testing. The tank and manifold were

constructed of Al 6061. All temperature readings on the thruster, including those on and in the nozzles were taken with

Omega type K thermocouples (range -200 °C - 1250 °C, error greater of 2.2 °C or 0.5%).

Silver mesh procured from fuelcellmaterials was used as the catalyst (nominal aperture 0.4 mm, wire diameter 0.11

mm). For all experiments presented in this study seven sheets were compacted into the catalyst chamber against a

stainless steel perforated retaining sheet. Prior to loading, the catalyst was baked at 500 °C for >30 minutes. The H2 O2

solution was procured from PeroxyChem at a 94% H2 O2 by mass.

In this iteration two major changes were made to the manifold design from the Prototype 2 design presented in Reference

3. First, the flow path between the tank and nozzle was increased from 1.8 mm to 5.3 mm. Second, the two valves were

3
replaced with one all aluminum valve with Viton® seals. These design changes addressed two of the major concerns

from the previous iteration: pressure drop and corrosion.


Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 3 Hydrogen peroxide vapor thruster Prototype 3: test unit (left), CAD model (right).

1. Catalyst Chamber/ Nozzle

Two different nozzle mounting configurations were tested in this series: (1) the 3 screw, Kalrez® o-ring variant that was

used in Prototype 1 and 2 and (2) a 6 screw, Vespel® spacer variant intended to improve thermal isolation and seal

integrity. These will be referred to as the o-ring and spacer variant for the remainder of this paper.

The spacer design prevents the nozzle from direct contact with the aluminum manifold, using the Vespel’s low thermal

conductivity (0.35 W/m/°C at 40 °C, compared to stainless steel’s 16 W/m/°C) to limit heat transfer. The spacer design

also utilizes a symmetric screw pattern to decrease the potential for thermal cycle induced loosening. Figure 4 shows

the manifold, o-ring/seal, and nozzle for both of the variants, specifically showing the difference in the sealing, screw

holes, and isolation.

4
Fig. 4 Prototype 3 mounting configurations: o-ring (left), spacer (right).
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

The nozzle and catalyst chamber were also varied. The two catalyst chambers/nozzles tested were (1) a thin-walled SS

316 design where temperature was measured along the exterior of the nozzle and (2) a MACOR® rectangular cuboid

with the capability to take 3 interior temperature measurements and 1 pressure measurement downstream of the catalyst.

Both nozzles featured the same interior converging diverging (CV) design with an interior catalyst chamber 5 mm in

diameter and length. The nozzle throat diameter was 0.79 mm and the exit diameter was 4.8 mm (area ratio 36.6). These

will be referred to as the SSCD nozzle and the MCD nozzle for the remainder of this paper. A close up of the SSCD and

MCD nozzles is shown in Figure 5.

Fig. 5 SSCD nozzle (left), MCD nozzle (right).

Thrust measurements were made with all configurations. For the o-ring variant, measurements were made for both the

MCD and the SSCD nozzles. For the spacer variant, measurements were made for a SSCD nozzle and a SS converging

only nozzle, which will be referenced as the SSCO nozzle. All interior dimensions and wall thicknesses were designed

to be the same for the spacer and o-ring configurations, however due to the complex machining required for the small

nozzles some minor variations in throat and exit diameter are noted in Table 1. Id. A, B, etc. will be used to reference

the various builds. Models of the nozzles used in the spacer variant can be seen in Figure 6.

5
Fig. 6 Spacer variant nozzles: SSCD (left), SSCO (right).

Table 1 Measured Nozzle Throat and Exit Diameters

Nozzle Throat Exit


Id.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Characteristic Diameter (mm) Diameter (mm)


A O-ring MCD 0.7 ± 0.013 4.6 ± 0.1
B O-ring SSCD 0.6 ± 0.013 6.0 ± 0.1
C Spacer SSCD 0.72 ± 0.013 5.5 ± 0.1
D Spacer SSCO 1 0.67 ± 0.013 -
E Spacer SSCO 2 0.67 ± 0.013 -

C. Thrust Measurement Setup

Thrust measurements were made with a torsional thrust stand consisting of an aluminum arm balanced on a frictionless

pivot with a calibrated spring constant. The thruster was placed on one side of the arm; for balance, counterweights were

placed on the other side. When the thruster was fired, the arm displacement was measured with an optical displacement

meter. Thrust was then calculated using the measured displacement and the spring constant,

𝜃𝑘
𝐹𝑇 = (2)
𝐿

where 𝐹𝑇 is thrust, 𝜃 is the deflection of the arm, 𝑘 is the spring constant, and 𝐿 is the moment arm.

The spring constant was experimentally determined using calibration electrodes to exert a known electrostatic force on

the test stand. This calibration was performed with the thruster mounted and the stand leveled and balanced, once at the

beginning of a test day and again at the end.

Figure 7 shows an image of the Prototype 3 thruster on the thrust stand. Further details on the design and calibration of

the specific thrust stand used in this study can be found in Hsu et al. [4].

6
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 7 Prototype 3 thruster on torsional thrust stand.

D. Results

Thrust, catalyst temperature, tank temperature, and tank pressure were measured (catalyst chamber pressure was also

measured for the MCD nozzle cases). Both the o-ring and the spacer mounting configurations were tested. All H2 O2 test

series were completed at tank temperatures of 60 °C, 70 °C, and 80 °C for varying run times (1, 5, and 10 minutes). The

two spacer variant nozzles were also tested with water (H2 O) as the propellant for a direct comparison of performance.

Water tests were performed at 30 °C, 40 °C, 50 °C, and 60 °C for the same run times. A separate test series was

completed for all of the mounting configurations, nozzles, and propellants to experimentally determine mass flow

through the system.

1. O-ring Variant

Two nozzles, the MCD and SSCD, were tested with the same o-ring mounting configuration as Prototype 2.

MCD Nozzle [Id. A]

For each of the three tank temperatures of interest (60 °C, 70 °C, and 80 °C) a series including six 1 minute tests, three

5 minute tests, and one 10 minute test was performed. The series was then completed with an additional test at 80

°C until all of the propellant was expelled. Catalyst temperature measurements were taken at 3 locations along the

centerline of the catalyst chamber in the MCD nozzle: (1) under the first sheet of silver mesh (top), (2) under the fourth

sheet (middle), and (3) under the final sheet (bottom), same as was done for the Prototype 2 test series. These catalyst

temperatures are shown for a 80 °C tank temperature test run featuring three 1 minute tests and two 5 minute tests in

7
Figure 8. The average catalyst temperature was taken as the mean of these measurements. Figure 9 shows the tank

pressure and chamber pressure for that same test run. Table 2 lists the steady state tank pressure, steady state pressure

drop to the catalyst chamber, and max average catalyst temperature for all of the MCD nozzle tests. The system pressure

drop is approximately 5% of the tank pressure, which is a substantial improvement over the previous iteration.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 8 Top, middle, and bottom catalyst temperatures in the O-ring MCD nozzle with 7 sheets silver mesh and
a tank temperature of 80 °C during a thrust measurement test series.

Fig. 9 Tank and catalyst chamber pressure in the Spacer MCD nozzle with 7 sheets silver mesh and a tank
temperature of 80 °C during a thrust measurement test series.

8
Table 2 Steady State Tank and Chamber Pressures and Maximum Average Catalyst Temperature for O-ring
MCD Nozzle Tests

Tank Steady State System Pressure Max Average Catalyst


Temperature (°C) Tank Pressure (Torr) Drop (Torr) Temperature (°C)
60 21.0 (-2.6 +3.4) 0.9 (-0.3 +0.4) 241 (-2.6 +4.4)
70 34.2 (-3.5 +4.7) 1.6 (- 0.2 +0.2) 317 (-5.8 +9.0)
80 51.2 (-5.4 +7.3) 2.8 (-0.6 +0.7) 387 (-6.0 +6.1)

Thrust and average catalyst temperature for all of the tests are shown in Figures 10 and 11, respectively. Variation in the

thrust level at a specific tank temperature is due to vacuum distillation of the liquid propellant throughout the test series.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Distillation decreased the vapor pressure and therefore the thrust.

Fig. 10 Measured thrust for all O-ring MCD nozzle thrust measurement test runs.

9
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 11 Average catalyst temperature for all O-ring MCD nozzle thrust measurement test runs.

Due to the nature of the thrust measurement test series (i.e., the entire series was completed with one propellant load),

mass flow measurements were not possible. Therefore an additional test series was performed with the same nozzle

configuration and more targeted test runs to experimentally determine the relationship between tank pressure, propellant

load, and mass flow rate. For each test a known quantity of propellant was loaded, the tank was set to either 60 °C, 70

°C, or 80 °C, and the test was run until the propellant was fully expelled. Figure 12 shows the mass flow relationship

with tank pressure. This linear fit was used to determine the Isp of each test performed on the thrust stand. The average

thrust and Isp for each tank temperature and length are shown in Table 3. Using the mass flow rate relation and run

times, the overall propellant usage was calculated to be 10.5 ml. This value is within 1% of the measured loaded value.

Fig. 12 O-ring MCD nozzle mass flow rate calculated from known propellant load and test run length.

10
Table 3 Thrust and Isp for O-ring MCD Nozzle Tests

Test Steady State Thrust (mN) Steady State Isp (s)


60 °C, 1 min 0.75 (-0.02 +0.02) 55.9 (-7.6 +6.8)
60 °C, 5 min 0.71 (-0.01 +0.02) 54.7 (-4.2 +7.6)
60 °C, 10 min 0.73 62.9
70 °C, 1 min 1.35 (-0.07 +0.05) 61.1 (-4.5 +4.3)
70 °C, 5 min 1.23 (-0.02 +0.03) 59.6 (-2.7 +4.8)
70 °C, 10 min 1.25 64.9
80 °C, 1 min 2.13 (-0.08 +0.11) 63.5 (-2.5 +2.6)
80 °C, 5 min 1.94 (-0.03 +0.03) 62.0 (-2.0 +3.4)
80 °C, 10 min 1.88 64.9
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

80 °C, 11 min 1.81 63.0

SSCD Nozzle [Id. B]

The test series performed on the MCD nozzle was repeated for the SSCD nozzle: same tank temperatures (60 °C, 70

°C, and 80 °C) and run times (1, 5, and 10 minutes). The catalyst chamber temperature measurement was taken on

the outside of the SS body at the throat. Due to higher temperatures at the o-ring in this design, leakage between the

manifold and nozzle prevented a fully successful test series. Figure 13 shows an example of a leak in the middle of a test

run, where the catalyst temperature drops off immediately with the thrust following shortly after. Any test with this

behavior was eliminated from the aggregated data of Table 4. Figure 15 and 14 show the catalyst temperatures and

thrust levels for all of the test runs without apparent leaks. However, major variations in temperature and thrust at the

same tank temperature suggests small leaks might have been prevalent throughout the series.

Fig. 13 Evidence of leak between manifold and catalyst chamber in 80 °C tank temperature run of O-ring
SSCD nozzle.

11
Table 4 H2 O2 Steady State Tank Pressures, Maximum Average Catalyst Temperature, Steady State Thrust,
and Steady State Isp for O-Ring SSCD Nozzle Tests

Tank Tank Press. Max Cat.


Thrust (mN) Isp (s)
Temp. (°C) (Torr) Temp. (°C)
60 21.9 (-2.2 +2.8) 195 (-11.1 +11.7) 0.57 (-0.05 +0.07) 58.2 (-9.1 +6.3)
70 36.2 (-6.0 +4.1) 255 (-22.4 +28.7) 1.07 (-0.13 +0.13) 66.2 (-6.5 +5.0)
80 62.1 (-3.1 +4.4) 327 (-17.9 +25.4) 1.88 (-0.11 +0.22) 67.4 (-3.6 +3.5)
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 14 Measured thrust for all O-ring SSCD nozzle thrust measurement test runs.

Fig. 15 Catalyst temperature for all O-ring SSCD nozzle thrust measurement test runs.

12
2. Spacer Variant

Two nozzles, the SSCD and the SSCO, were tested with the spacer mounting configuration. These nozzles were tested

both with H2 O2 as well as with H2 O to get a direct comparison of performance. In the H2 O tests, the silver catalyst was

removed.

SSCD Nozzle [Id. C]

For each of the 60 °C, 70 °C, and 80 °C tank temperatures, three 1 minute, two 5 minute and one 10 minute tests were

performed with H2 O2 as the propellant. Due to H2 O’s higher vapor pressure, H2 O test runs were performed at 30 °C,
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

40 °C, 50 °C, and 60 °C.

Temperature was measured on the outside of the SS catalyst chamber in 2 locations (1) immediately upstream of the

converging nozzle section and (2) at the throat. In both nozzles using the spacer mounting variant, temperature was

measured using a type K thermocouple with a SS sheath. An image is shown in Figure 16. Temperature measurements

for both thermocouples are shown for an H2 O2 , 60 °C tank temperature test series in Figure 17. Chamber temperatures

for the water tests are not shown, since they match the tank temperature within 1 °C.

Fig. 16 Thermocouple placement on Spacer SSCD nozzle.

Thrust and catalyst temperatures for the H2 O2 tests are shown in Figure 18 and 20, respectively. Catalyst temperature is

taken as the throat temperature measurement, for congruity with previous tests. The average steady state tank pressure,

maximum catalyst temperatures, steady state thrust, and steady state Isp for each of the tank temperatures are provided in

Table 5. Evidence of minor catalyst chamber leaking can be seen in the catalyst temperature and thrust during the 80 °C

tests.

Thrust measurements for the H2 O tests can be seen in Figure 19, with temperatures, pressures, thrust, and Isp provided

in Table 6. The system heaters and thermocouple feedback loop were designed for the lower evaporation rate associated

13
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 17 Temperature measured at the bottom of the catalyst chamber and throat for a 60 °C tank experiment
series on the Spacer SSCD nozzle running H2 O2 . Thermocouple placement shown in Figure 16.

with H2 O2 . This resulted in more tank temperature variation and therefore tank pressure variation with the H2 O tests.

This presents itself as a slight wobble in the thrust measurement. Small irregularities in the water evaporation also

appear in the pressure and thrust measurements, potentially due to lack of adequate nucleation sites.

Fig. 18 Measured thrust for all H2 O2 Spacer SSCD nozzle thrust measurement test runs.

14
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 19 Measured thrust for all H2 O Spacer SSCD nozzle thrust measurement test runs.

Fig. 20 Catalyst temperature for all H2 O2 Spacer SSCD nozzle thrust measurement test runs.

Table 5 H2 O2 Steady State Tank Pressures, Maximum Average Catalyst Temperature, Steady State Thrust,
and Steady State Isp for Spacer SSCD Nozzle Tests

Tank Tank Press. Max Cat.


Thrust (mN) Isp (s)
Temp. (°C) (Torr) Temp. (°C)
60 25.0 (-1.7 +1.3) 263 (-18.3 +7.7) 1.17 (-0.05 +0.05) 71.7 (-1.1 +2.1)
70 43.1 (-2.6 +2.6) 336 (-2.0 +1.1) 2.07 (-0.13 +0.16) 73.3 (-0.6 +1.5)
80 69.4 (-6.5 +5.3) 388 (-36.8 +9.8) 3.27 (-0.43 +0.32) 71.9 (-2.9 +1.6)

15
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Table 6 H2 O Steady State Tank Pressures, Maximum Average Catalyst Temperature, Steady State Thrust,
and Steady State Isp for Spacer SSCD Nozzle Tests

Tank Tank Press. Max Cat.


Thrust (mN) Isp (s)
Temp. (°C) (Torr) Temp. (°C)
30 29.6 (-0.5 +1.3) 30 (-0.3 +0.3) 1.31 (-0.03 +0.05) 48.6 (-1.0 +1.7)
40 54.0 (-2.0 +1.8) 40 (-0.7 +0.6) 2.45 (-0.10 +0.09) 50.1 (-0.2 +0.3)
50 87.8 (-6.8 +4.9) 50 (-0.5 +0.2) 4.13 (-0.40 +0.36) 51.8 (-1.1 +1.6)
60 136.7 (-1.6 +1.6) 59 (-0.1 +0.1) 6.73 (-0.29 +0.29) 54.3 (-1.7 +1.7)

16
SSCO Nozzle [Id. D and E]

The H2 O2 and H2 O tests performed on the spacer SSCD nozzle were repeated with a SSCO nozzle. Due to the lack of a

diverging section, temperature could only be taken on the bottom exterior of the catalyst chamber. However, as was

shown in Figure 17, this temperature closely matches that of the throat. For the SSCD test series, both the H2 O2 and

H2 O tests were performed with the same nozzle build (Id. C, catalyst removed for water tests). In this test series, with

the SSCO configuration, the H2 O2 tests and H2 O tests were performed with different nozzle builds, Id. D and Id. E,

respectively. While they have the same measured throat diameter, as detailed in Table 1, there appeared to be a burr in

the flow path of the nozzle used in the water test series, which decreased the mass flow rate for that test series.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Thrust for the H2 O2 and H2 O tests are shown in Figure 21 and Figure 22, respectively. The catalyst temperatures for

the H2 O2 test are shown in Figure 23. Averaged data are provided in Table 7 and Table 8. For the H2 O2 tests, leaks

between the catalyst chamber and manifold are evident during the 80 ° tank test series; this shows up in both the thrust

measurement and in the catalyst temperatures. Both the H2 O2 and H2 O thrust measurements show some variance in tank

temperature/pressure/thrust. Improper heater placement/control resulted in more dramatic swings in tank temperature.

Also the H2 O SSCO tests show small irregular pops in pressure and thrust, as was seen in the previous H2 O SSCD test

series.

Fig. 21 Measured thrust for all H2 O2 Spacer SSCO nozzle thrust measurement test runs.

17
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 22 Measured thrust for all H2 O Spacer SSCO nozzle thrust measurement test runs.

Fig. 23 Catalyst temperature for all H2 O2 Spacer SSCO nozzle thrust measurement test runs.

Table 7 H2 O2 Steady State Tank Pressures, Maximum Average Catalyst Temperature, Steady State Thrust,
and Steady State Isp for Spacer SSCO Nozzle Tests

Tank Tank Press. Max Cat.


Thrust (mN) Isp (s)
Temp. (°C) (Torr) Temp. (°C)
60 25.7 (-1.0 +0.9) 232 (-5.6 +4.0) 1.01 (-0.02 +0.02) 68.7 (-2.1 +2.3)
70 42.9 (-2.6 +1.4) 309 (-5.9 +5.4) 1.74 (-0.09 +0.05) 70.7 (-0.9 +0.5)
80 66.9 (-4.8 +4.2) 345 (-42.8 +29.7) 2.63 (-0.26 +0.18) 68.5 (-1.9 +1.7)

18
Table 8 H2 O Steady State Tank Pressures, Maximum Average Catalyst Temperature, Steady State Thrust,
and Steady State Isp for Spacer SSCO Nozzle Tests

Tank Tank Press. Max Cat.


Thrust (mN) Isp (s)
Temp. (°C) (Torr) Temp. (°C)
30 30.7 (-0.3 +0.9) 30 (-0.2 +0.2) 0.86 (-0.03 +0.02) 52.6 (-1.2 +1.6)
40 55.5 (-1.4 +1.6) 40 (-0.3 +0.7) 1.56 (-0.05 +0.05) 53.0 (-0.5 +0.4)
50 92.0 (-3.1 +3.1) 49 (-0.3 +0.5) 2.53 (-0.08 +0.11) 51.9 (-0.6 +0.5)
60 139.0 (-8.2 +6.4) 58 (-0.5 +0.7) 3.89 (-0.24 +0.18) 52.8 (-0.5 +0.6)
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

E. Analysis

Several factors influence the thrust and overall system performance. In order to directly compare the different

configurations, thrust and Isp for all of the cases were normalized to the o-ring MCD nozzle mass flow rate at the 60 °C,

70 °C, and 80 °C (corresponding to 30 °C, 40 °C, and 50 °C for H2 O). This adjustment compensated for the different

throat diameters. A direct comparison of the thrust and Isp for all of the Prototype 3 designs and propellants is provided

in Figure 24 and Figure 25, respectively. The spacer variant SSCD nozzle operating with H2 O2 propellant demonstrated

the highest thrust and Isp , approximately 60% higher than the same nozzle operating on H2 O.

Fig. 24 Average normalized thrust for all Prototype 3 thrust measurement tests.

19
Fig. 25 Average normalized Isp for all Prototype 3 thrust measurement tests.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

1. Comparison with Theoretical Nozzle

A chemical equilibrium code based on STANJAN [5] was used to compute the isentropic mass flow rate, thrust, and

vacuum specific impulse for comparison with the spacer variant nozzle experimental values; results are shown in Table

9. Throat Reynolds number, 𝑅𝑒 𝑡 , and Knudsen number, 𝐾𝑛𝑡 , were also computed using Equations 3, 4, and 5. The

input parameters of the analyses were the average measured tank pressures and average maximum catalyst chamber

temperatures. The analysis assumed complete decomposition and isentropic expansion of the products by a specified

exit-to-throat area ratio. From the pressure, velocity, and density at the throat (defined as the location where the local

sound speed is equal to the local velocity, i.e., a Mach number of unity) and the exit were calculated. These calculations

assume continuum flow and no boundary layer effects.

𝑢 𝑡 𝑑𝑡
𝑅𝑒 𝑡 = (3)
𝜈𝑡
𝑘 𝐵 𝑇0
𝜆= √ (4)
2𝜋𝑑 2 𝑃0
𝜆
𝐾𝑛𝑡 = (5)
𝑑𝑡

where 𝑢 𝑡 is the throat flow speed, 𝑑𝑡 is the throat diameter, 𝜈𝑡 is the throat kinematic viscosity, 𝜆 is the gas mean free

path, 𝑘 𝐵 is the Boltzmann constant, 𝑇0 is the stagnation temperature or catalyst temperature, 𝑑 is the average molecular

diameter, and 𝑝 0 is the stagnation pressure or chamber pressure.

The mass flow rate ratio, which is often referred to as discharge coefficient, 𝐶𝐷 , appears directly correlated with

Reynolds number. Higher Reynolds number leads to a 𝐶𝐷 approaching unity. Note the SSCO nozzle with H2 O as the

propellant is disregarded in this analysis. Post-testing a large burr was found in the throat region, which would have a

20
Table 9 Comparison of the Spacer Variant Nozzle Experimental Mass Flow Rate, Thrust, and Isp with that of
an Isentropic Nozzle in Continuum Flow

Tank
Nozzle m¤ Thrust Isp
Prop. Temp. Ret Knt
Type/ Id. (Ex./Th.) (Ex./Th.) (Ex./Th.)
(°C)
SSCD/ C H 2 O2 60 171 0.0084 0.77 0.51 0.67
SSCD/ C H 2 O2 70 245 0.0059 0.83 0.53 0.64
SSCD/ C H 2 O2 80 344 0.0042 0.87 0.52 0.60
SSCO/ D H2 O2 60 176 0.0082 0.78 0.70 0.90
SSCO/ D H2 O2 70 240 0.0060 0.83 0.73 0.87
SSCO/ D H2 O2 80 343 0.0042 0.86 0.70 0.82
SSCD/ C H2 O 30 553 0.0026 0.97 0.49 0.51
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

SSCD/ C H2 O 40 955 0.0015 0.99 0.51 0.51


SSCD/ C H2 O 50 1479 0.00098 1.00 0.52 0.52
SSCO/ E H2 O 30 533 0.0027 0.64 0.5 0.78
SSCO/ E H2 O 40 915 0.0016 0.65 0.50 0.77
SSCO/ E H2 O 50 1441 0.0010 0.66 0.49 0.74

dramatic effect on flow characteristics, specifically mass flow rate.

This relationship between 𝑅𝑒 and 𝐶𝐷 has been established by other researchers and the experimental values determined

here match relatively well with the relation developed by Kuluva et al. [6] (Equation 6). Experimental H2 O2 𝐶𝐷 values

were within 5% of those found via the equation, with the H2 O 𝐶𝐷 values within 11%.

  0.019 "   0.21   0.5 #


𝑟 𝑐 + 0.05𝑟 𝑡 𝑟 𝑐 + 0.10𝑟 𝑡 1
𝐶𝐷 = 1− 𝑓 (𝛾) (6)
𝑟 𝑐 + 0.75𝑟 𝑡 𝑟𝑡 𝑅𝑒

𝑓 (𝛾) ≈ 0.97 + 0.86𝛾 (7)

where 𝑟 𝑐 is the radius of the chamber, 𝑟 𝑡 is the radius of the throat, and 𝛾 is the specific heat ratio.

The discrepancy in the theoretical and experimental thrust and Isp is not completely explained by the differences in mass

flow rates, especially for the nozzles with a diverging section. However with a 𝑅𝑒 < 1000 for a majority of the tests,

viscosity losses are a known dominate factor in nozzle performance. Inefficiencies arise from the adverse interaction of

the subsonic boundary layer near the wall with the supersonic flow at the core. For low Reynolds numbers the viscous

boundary layer can occupy almost all of the diverging section. Bruccoleri et al. [7] found nozzle efficiencies of around

75% for 𝑅𝑒 > 1500, but that dropped to approximately 60% for 𝑅𝑒 ≈ 400, and 40% for 𝑅𝑒 < 200. Murch et al. [8] also

found that the inviscid core/ boundary layer interaction was an important effect that scaled with Reynolds number. Kirn

21
[9] studied the viscous and divergence losses for nozzles with Reynolds numbers ranging from 270 to 1150 using a full

Navier Stokes code. Results showed large viscous boundary layers, which dominate an increasing portion of the exit

plane with increasing nozzle length.

In addition to viscous losses, divergence losses also decrease performance with respect to theory. The SSCD design

features an especially large nozzle angle, 𝛼, of 67°. Thrust loss due to divergence is typically treated by

¤ 𝑒 + ( 𝑝 𝑒 − 𝑝 𝑎 ) 𝐴𝑒
𝐹𝑐𝑜𝑛𝑖𝑐,𝑎 𝑝 𝑝𝑜𝑥 = 𝜆 𝑚𝑢 (8)

1 + cos 𝛼
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

𝜆= (9)
2

where 𝐹𝑐𝑜𝑛𝑖𝑐,𝑎 𝑝 𝑝𝑜𝑥 is the theoretical thrust approximation for a conical nozzle with a nozzle angle of 𝛼.

Incorporating this into the isentropic theoretical calculations for the SSCD nozzle deceases the potential thrust by

approximately 30%. The converging only designs, SSCO, would also experience divergence losses due to flow expansion

at the exit plane.

Due to the low Reynolds numbers and Knudsen numbers bordering the transitional flow regime (0.01 < 𝐾𝑛 < 100),

theoretical calculations of mass flow rate, thrust, and Isp in a free molecular flow were also completed:

𝑝 0 𝐴𝑡
𝐹𝑇 ,𝐹 𝑀 = (10)
2
r
𝜋 𝑘𝐵 1
𝐼𝑠 𝑝,𝐹 𝑀 = 𝑇0 (11)
2 𝑀 𝑔
𝐹𝑇 ,𝐹 𝑀
𝑚¤ 𝐹 𝑀 = (12)
𝐼𝑠 𝑝,𝐹 𝑀 𝑔

where 𝐹𝑇 ,𝐹 𝑀 is the free molecular thrust, 𝐴𝑡 is the throat area, 𝑀 is the molecular mass, and 𝑔 is the gravitational

constant.

Ratios of the experimental mass flow, thrust, and Isp to that of free molecular flow theory are provided in Table 10.

Based on the lack of difference in thrust and Isp ratios with Knudsen number, all the experiments performed in this study

appear to fall in the continuum regime. In the free molecular or transitional regime, higher Knudsen numbers would

have ratios approaching unity.

22
Table 10 Comparison of the Spacer Variant Nozzle Experimental Mass Flow Rate, Thrust, and Isp with that
of an Isentropic Nozzle in Free Molecular Flow

Tank
Nozzle m¤ Thrust Isp
Prop. Temp. Ret Knt
Type/ Id. (Ex./Th.) (Ex./Th.) (Ex./Th.)
(°C)
SSCD/ C H 2 O2 60 171 0.0084 1.24 1.72 1.39
SSCD/ C H 2 O2 70 245 0.0059 1.32 1.77 1.34
SSCD/ C H 2 O2 80 344 0.0042 1.37 1.74 1.26
SSCO/ D H2 O2 60 176 0.0082 1.24 1.67 1.35
SSCO/ D H2 O2 70 240 0.0060 1.33 1.73 1.30
SSCO/ D H2 O2 80 343 0.0042 1.37 1.67 1.22
SSCD/ C H2 O 30 553 0.0026 1.54 1.63 1.06
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

SSCD/ C H2 O 40 955 0.0015 1.57 1.67 1.07


SSCD/ C H2 O 50 1479 0.00098 1.59 1.73 1.09
SSCO/ E H2 O 30 533 0.0027 1.02 1.19 1.16
SSCO/ E H2 O 40 915 0.0016 1.04 1.20 1.15
SSCO/ E H2 O 50 1441 0.0010 1.06 1.17 1.11

F. Comparison with Theoretical Catalyst Temperature

Average catalyst temperatures for all of the Prototype 2 and 3 nozzles, with H2 O2 as the propellant, are shown in Figure

26. When comparing the Prototype 2 nozzles, which were detailed in Reference 3, to that of their equivalent Prototype

3 nozzles, there is an improvement of approximately 30 - 40% depending on the tank temperature. However even the

maximum temperature seen in these tests still falls well below the adiabatic temperatures.

Fig. 26 Average catalyst temperatures for all Prototype 2 and 3 thrust measurement tests.

23
1. Heat Transfer Analysis using Finite Element Modeling

COMSOL was used to model the two mounting configurations and their respective nozzle. Due Prototype 3’s lower

system pressure drop (i.e., higher chamber pressure) and higher liquid H2 O2 concentration (i.e., higher vapor H2 O2

concentration), the heat flux values used to model the decomposition process were slightly higher than the Prototype 2

values at 2.65 W, 4.32 W, and 6.47 W for tank temperatures of 60 °C, 70 °C, and 80 °C, respectively.

Table 11 compares the average catalyst temperatures for the Prototype 2 nozzles and those of Prototype 3, both the o-ring

and spacer variants. Images of the Prototype 3 COMSOL models sectioned along the centerline are shown in Figures

27 and 28. Overall the model trends for each nozzle style are reflected in the experimental values shown in Figure
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

26. Measured and modeled temperatures increased for the MCD nozzle from Prototype 2 to Prototype 3. Measured

and modeled temperatures also increased for the SSCD nozzle from Prototype 2 to the o-ring variant of Prototype 3

to the space variant of Prototype 3. One deviation was with the spacer SSCO nozzle; the model predicted that the

catalyst temperatures would be higher than the spacer SSCD nozzle, which was not realized in the thrust measurement

experimental series.

24
Table 11 COMSOL Catalyst Temperatures for Prototype 2 and 3

Prototype 2 Prototype 3
O-Ring O-Ring Spacer
Tank MCD SSCD MCD SSCD SSCD SSCO
Temp. Nozzle Nozzle Nozzle Nozzle Nozzle Nozzle
60 °C 101 °C 116 °C 115 °C 132 °C 197 °C 227 °C
70 °C 142 °C 163 °C 161 °C 184 °C 281 °C 320 °C
80 °C 199 °C 224 °C 217 °C 243 °C 375 °C 421 °C
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 27 Kalrez® O-Ring variant COMSOL model slice temperature for 80 °C tank temperature: MCD nozzle
(left), SSCD nozzle (right).

25
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Fig. 28 Vespel® Spacer variant COMSOL model slice temperature for 80 °C tank temperature: SSCD nozzle
(left), SSCO nozzle (right).

II. Conclusions

The development and test of Prototype 2 allowed for the investigation of 3 different catalyst materials and the identification

of the highest performing material and bed length: 7 sheets of silver mesh with an aperture of 0.42 mm and a wire

diameter of 0.11 mm. Measured catalyst temperatures were approximately 30% of their adiabatic temperature, assuming

full decomposition. A finite element model of the system heat transfer was developed to identify and help inform design

modifications. The model showed that conduction from the nozzle to the manifold was the primary heat loss path.

Minimizing that loss could improve catalyst temperatures by up to 135%.

While valuable data was collected in the Prototype 2 series, high system pressure drop and mild valve corrosion spurred

the design of Prototype 3. In Prototype 3, the flow path diameter was increased and the valves were replaced with

a more compatible, larger orifice option. Using a torsional thrust stand, thrust measurements were made with the

Prototype 3 design with four different nozzle configurations: (1) a MACOR® converging diverging nozzle with a Viton®

o-ring seal, (2) a stainless steel converging diverging nozzle with a Viton® o-ring seal, (3) a stainless steel converging

diverging nozzle with a Vespel® spacer seal, and (4) a stainless steel converging only nozzle with a Vespel® spacer seal.

The last two nozzle configurations were tested with both H2 O2 and H2 O as the propellant, for a direct comparison of

performance. Testing revealed that the Spacer SSCD nozzle with H2 O2 was the highest performer in terms of both

thrust and specific impulse. It outperformed its H2 O counterpart by 60%.

Comparison of the experimental thrust results with a theoretical isentropic, inviscid nozzle showed the dramatic

impact of the low Reynolds number viscous boundary layer (ReH2 O2 < 350, ReH2 O < 1500). Thrust and Isp ratios

26
(experimental/theoretical) varied from 0.49 - 0.73 and 0.51 - 0.9, respectively.

A 30 - 40% increase in catalyst temperature was realized between the Prototype 2 and 3 configuration, which showed

the benefit of the larger manifold flow area and the additional thermal isolation. A heat transfer model of the Prototype 3

design was compared to that of Prototype 2; the models matched the experimental trends.

Future iterations will focus on increasing thermal isolation and diverging nozzle design. Better isolation of the

catalyst chamber from the body could further increase the catalyst temperatures and therefore the overall performance.

Redesigning the nozzle for low Reynolds number flow could also increase performance by minimizing the subsonic

boundary layer.
Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

Acknowledgments

This research was supported by The Aerospace Corporation’s Independent Research and Development program.

References

[1] Cervone, A., Torre, L., D’Agostino, L., Musker, A. J., Roberts, G. T., Bramanti, C., and Saccoccia, G., “Development

of Hydrogen Peroxide Monopropellant Rockets,” AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, 2006.

doi:10.2514/6.2006-5239.

[2] Rhodes, B. L., and Ronney, P. D., “Dynamics of a Small-Scale Hydrogen Peroxide Vapor Propulsion System,” Journal of

Propulsion and Power, Vol. 35, No. 3, 2019, pp. 595–600. doi:10.2514/1.B37323.

[3] Rhodes, B. L., Ulrich, E. R., and Ronney, P. D., “Small-Scale Hydrogen Peroxide Vapor Propulsion System: Catalyst

Performance and Heat Transfer,” AIAA Propulsion and Energy Forum, American Institute of Aeronautics and Astronautics

(AIAA), Indianapolis, IN, 2019. doi:10.2514/6.2019-4029.

[4] Schouten, A. G. H., Beiting, E. J., and Curtiss, T. J., “Performance of a Torsional Thrust Stand with 1 𝜇N Sensitivity,” Joint

Conference of 30th International Symposium on Space Technology and Science 34th International Electric Propulsion Conference

and 6th Nano-satellite Symposium, Kobe-Hyogo, Japan, 2015.

[5] Reynolds, W. C., “The Element Potential Method for Chemical Equilibrium Analysis : Implementation in the Interactive Program

STANJAN, Version 3,” Tech. rep., Dept. of Mechanical Engineering, Stanford University, Stanford, 1986.

[6] Kuluva, N. M., and Hosack, G. A., “Supersonic Nozzle Discharge Coefficients at Low Reynolds Numbers,” AIAA Journal, Vol. 9,

No. 9, 1971, pp. 1876–1879. doi:10.2514/3.6443.

27
[7] Bruccoleri, A. R., Leiter, R., Drela, M., and Lozano, P., “Experimental Effects of Nozzle Geometry on Flow Efficiency at Low

Reynolds Numbers,” Journal of Propulsion and Power, Vol. 28, No. 1, 2012, pp. 96–105. doi:10.2514/1.B34073.

[8] Murch, C. K., Broadwell, J. E., Silver, A. H., and Marcisz, T. J., “Performance Losses in Low-Reynolds-Number Nozzles,”

Journal of Spacecraft and Rockets, Vol. 5, No. 9, 1968, pp. 1090–1094. doi:10.2514/3.29426.

[9] Kirn, S. C., “Calculations of Low-Reynolds-Number Resistojet Nozzles,” Journal of Spacecraft and Rockets, Vol. 31, No. 2,

1994, pp. 259–264. doi:10.2514/3.26431.


Downloaded by CORNELL UNIVERSITY on August 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3812

28

You might also like