You are on page 1of 25

Measurement of rocket performance variation with propellant mixture and preparation

Kieran F. Dineen1

This thesis report aims to investigate the sensitivity of homemade rocket propellant performance to variables such as chemical composition, particle size and grain shape. This will be achieved by measuring the thrust-time relationship for each possible combination within the time constraint by using a test rig manufactured by Morgan Carter in 2008. This type of analysis could prove valuable in any area of solid propellants where the need for more powerful, lighter, cheaper safer and more environmentally friendly propellant is desired. The thesis will initially focus on performance of several homemade motors followed by comparison of thrust-time relationships to those of the commercially available Estes motors and then conclude with optimisation of a homemade propellant mixture.

Aeronautical Engineering: Project, Thesis and Practical Experience ZACM4049/4050 1

Final Thesis Report 2009, ACME, UNSW@ADFA

Table of Contents
Measurement of rocket performance variation with propellant mixture and preparation Nomenclature I. II. A. B. 1. 2. 3. C. 1. 2. 3. D. III. IV. V. VI. A. B. C. D. E. VII. A. B. C. D. E. F. G. VIII. A. B. IX. A. B. C. X. Introduction Background Introduction Rocket principles Thrust Impulse Mass ratio and propellant mass fraction Propellant Burning rate Grain configuration Chemical composition Measurement Thesis methodology Commercially available motors Simulation Homemade rocket construction Rocket cases Igniters Aluminium fuel propellant Sucrose fuel propellant Packing the motor cases Experiment method Cantilever beam Item 1 Amplifier Item 2 Laptop/USB DAQ Item 3 Digital scales Item 4 Safety Perimeter Item 5 Ignition system Item 6 Experiment procedure Results Aluminium fuel propellant Sucrose fuel propellant Discussion Performance of Aluminium based propellant Low thrust and load cell resolution Thrust peaks and nozzle failure Conclusion 1 3 4 4 4 5 5 5 6 6 6 7 7 9 10 11 11 13 13 13 14 14 15 15 16 16 16 16 17 17 17 17 17 18 19 19 19 21 22

2 Final Thesis Report 2009, ACME, UNSW@ADFA

XI.

Recommendation

22 22 23 25

Acknowledgements References Appendices

Nomenclature
F 2 p1 p2 p3 A2 g0 tb It Is FT mp mf m0 Ab r n a b Kn M R T1 c* = = = = = = = = = = = = = = = = = = = = = = = = = = momentum thrust force [N] mass flow rate of propellant [kg/s] instantaneous exhaust velocity [m/s] internal pressure of rocket motor [Pa] pressure at motor nozzle exit [Pa] ambient pressure [Pa] nozzle exit area [m2] acceleration due to gravity [m/s] burn rate of propellant [s] total impulse [N s] specific impulse [s] total thrust force [N] mass of propellant [kg] final mass of rocket [kg] initial mass of rocket [kg] burning area [m2] burn rate [m/s] burn rate exponent temperature coefficient [m/s] propellant density [kg/m] Ab/At molecular weight of products [kg/ kmol] specific gas constant [J/ mol.K] chamber temperature [K] characteristic propellant velocity [m/s] ratio of specific heats

3 Final Thesis Report 2009, ACME, UNSW@ADFA

I.

Introduction

HIS thesis represents ongoing work in the area of solid rocket propellant science and engineering at UNSW@ADFA. Initial work conducted by Morgan Carter in 2008 investigated performance of commercially available solid rocket motors manufactured by Estes Industries. This thesis follows the recommendations of Carter and continues investigation into homemade solid rocket propellants. Solid rockets have many applications from hobby rockets to launching massive payloads into space. They are often used in SCRAMJET research due to the ability to easily produce high velocity conditions. Military uses range from air to air missiles to shoulder launched weapon platforms. There are many other applications such as ejection systems on aircraft, assist-take-off rockets for airplanes, target drone propulsion and signal flares (Sutton, Biblarz 2001, p25). Amateur rocketry is also another use for small solid rocket motors. In the mid 1950s when the space age began flying rockets became very popular. At the time there were no safe propellants readily available to launch rockets. In 1958 Vernon Estes developed the first safe, mass-produced model rocket engine (Estes Catalogue, 2008, p4). Although commercially available rocket motors are safe and reliable they are quite expensive, costing as much as US$18.90 for a pack of three motors (Estes Catalogue, 2008, p4). Launch costs could be significantly reduced if a reliable motor with similar performance could be made safely from readily available substances.

II.

Background

A. Introduction Rockets operate on the principle of Newtons third law of motion that for every action there is an equal and opposite reaction. The combustion of propellant results in formation of hot gases which are accelerated at high velocity through a nozzle imparting momentum to the engine (Braeunig). In solid rockets the propellant refers to the mixture of fuel and oxidiser within an engine. The propellant is molded into a solid shape inside the rocket motor which is called the propellant grain (Sleeter 2004, p20). On ignition a chemical reaction occurs between the fuel and oxidiser which produces the hot gases required to produce thrust. Once the grain is ignited thrust cannot be throttled or stopped. The grain will burn at a set rate until all propellant is consumed. Thrust produced from a rocket engine can range from as little as 4 N to over 4 million N (Sutton, Biblarz 2001, p417). To produce this thrust rocket motors can produce temperatures up to 4100C with exhaust velocities from the nozzle reaching 4300 m/s (Sutton, Biblarz 2001). Almost all rockets are used only once. In rare cases such as the Shuttle solid booster the hardware is recovered and refurbished for further use. The added cost of designing such device would only result in savings if the device was used often enough. For the design of the homemade rocket it would be desirable to have a reusable motor case as this could significantly reduce costs. The similarities between large scale rocket motors and commercially available hobby rocket motors can be seen by comparing Figure 1 and Figure 2.

Figure 1. Rocket motor cross section showing typical design (Purdue University). This figure shows components present in most rocket motors.

Figure 2. Commercially available Estes rocket motor cross section (Estes Catalogue, 2008, p38). This figure shows components present in the commercially available Estes brand rocket motors.

4 Final Thesis Report 2009, ACME, UNSW@ADFA

B. Rocket principles The principles used in rocket propulsion are those of mechanics, thermodynamics and chemistry. Propulsion is the act of applying a force to accelerate an object or maintain a constant velocity against a resistance. In a rocket this propulsive force is obtained by ejection of high velocity gas through a nozzle. 1. Thrust A simplified view of thrust is that it is the reaction experience by the rocket structure due to the ejection of matter at a high velocity. This ejection of matter imparts momentum to the rocket structure. The equation giving thrust, caused by change in momentum is given below (Sutton, Biblarz 2001, p32): =

2 = 2

(1)

This force, F (N) represents thrust when the pressure at the nozzle exit is equal to the ambient pressure. This makes the assumption that thrust and mass flow are constant and velocity is uniform and axial. The second type of thrust is known as pressure thrust and arises when there is a pressure difference between the nozzle exit and the surrounding atmosphere. From this the total thrust FT (N) is given by the following equation (Sutton, Biblarz 2001, p32): = 2 + (2 3 )2 (2)

The second term on the right hand side of Eq. (2) is pressure thrust. Pressure thrust is the product of the nozzle exit area A2 (m2) and the difference between the nozzle exit pressure, p 2 (Pa) and surrounding pressure p3 (Pa). Figure 3 shows a schematic of the pressure acting on a rocket motor. Pressure thrust is determined by integrating all pressures normal to the nozzle plane. Due to the axially symmetric design of rockets the pressures acting radially outward do not contribute to thrust. For this reason the only values of concern are p2, p3 and A2. Positive thrust is produced if the nozzle exit pressure is greater than the ambient pressure (under-expanded condition, p2>p3). Conversely, if the ambient pressure is greater than that of the nozzle exit (over-expanded, p2<p3), negative pressure thrust is produced. The optimum Figure 3. Schematic diagram of pressure acting on a expansion ratio is when p2=p3 and only rocket motor (Sutton, Biblarz 2001, p33). Magnitude momentum thrust is produced. Therefore nozzles of arrows show the relative pressure, subscript 1, t, 2 and are designed to have either an optimum or under 3 denote conditions internal, at the throat, at nozzle exit expanded flow (Sutton, Biblarz 2001, p33). The and external to the rocket. geometry of the nozzles for the testing cases have been produced assuming the same chamber conditions as the Estes motors (Carter 2008, p28).

2. Impulse Total impulse It (Ns) is thrust force (which can vary over time) integrated over the burning time, tb (s). =
0

(3)

If transient forces at the start and end of the burn are considered negligible then this simply becomes. = (4)

It is proportional to the energy released by the rocket during a complete burn. A more useful quantity used in rocketry is specific impulse I s (s). This figure is a way to compare rocket performance much like cars are comparable by mileage; how many liters of fuel consumed per 100 kilometers.

5 Final Thesis Report 2009, ACME, UNSW@ADFA

Higher specific impulse generally means better performance although optimal values of specific impulse are dependent on the mission requirements. Specific impulse is the impulse per unit weight of propellant. =

(5)

Equation 5 will give a time averaged impulse taking into account the transient effect of starting and stopping of the motor. This can be simplified for constant mass flow and thrust for negligible star and stop times. For this case the mass is the effective propellant mass, mp (kg) resulting in Eq. (6) (Sutton, Biblarz 2001, p28). Specific impulse will be one parameter used to compare the different homemade rockets to the Estes motors =
0

(6)

3. Mass ratio and propellant mass fraction Mass ratio, MR is the ratio of the final mass of the rocket mf (kg) to the initial mass of the rocket m0 (kg) (sum of empty mass and propellant mass), shown in Eq. (7). This is another figure used to compare rocket performance. It will give an indication of how much of the propellant is used for a given mixture. Lower mass ratios indicate that more of the propellant is being used with a value close to one meaning very little propellant is being used. The lowest value is the ratio of the empty mass to full mass of the rocket (Sutton, Biblarz 2001, p29). =
0

(7)

C. Propellant The operation and design of a motor depend on the combustion characteristics of the propellant, its burning rate, burning surface and grain geometry. The branch of science describing these values is known as internal ballistics. It is these characteristics that will change for each variation in propellant formula. 1. Burning rate Once ignited the propellant grain will recede in the direction perpendicular to the surface. The burning rate r (cm/s) describes the rate of regression of the propellant grain. Burning rate depends on a number of factors including particle size, particle shape, grain configuration, chamber pressure and oxidiser percentage (Sutton, Biblarz 2001, p426). Burning rate is important as it determines the mass flow of the rocket which in turn determines thrust (refer to Eq. (1)). The relation between mass flow and burning rate is given by Eq. (8) where Ab (m2) is the burning area and b (kg/m3) is the propellant density (Sutton, Biblarz 2001, p427). = (8)

Analytical modeling and the supportive research has yet to predict the burning rate of a new propellant in a new motor. Elemental laws and equations on burning rate usually deal with the influence on some of the important parameters individually. Often it is possible to approximate burning rate as a function of chamber pressures over a limited range of chamber pressures. The relation is given by Eq. (9) and applies to commonly used double-base, composite, or composite double base propellants. = 1 (9)

In this equation a is known as the temperature coefficient and has units of m/s. It is influenced by the ambient temperature of the grain. The burning rate exponent n is independent of the grain temperature and describes the influence of chamber pressure on burning rate. High values of n give rapid change of burning rate with pressure which means that even a small change in pressure will result in a substantial change in hot gas produced (Sutton, Biblarz 2001, p428). It is important to note that if the propellant used is particularly sensitive to pressure increase, burning will spiral of control and could possibly lead to explosion. For this reason it is recommended that a propellant with a constant burning rate over a range of pressures is used (Brinley, Remington 1960, p77).

6 Final Thesis Report 2009, ACME, UNSW@ADFA

As mentioned previously, particle size and shape has an influence on the burning rate. Spherical shaped particles allow for easier mixing and higher percentage of solids in the propellant. Smaller size particles maximize oxidiser per unit volume of propellant by filling in smaller voids between larger particles (Sutton, Biblarz 2001, p503). The effect of particle size on burning rate can be seen in Figure 4. The temperature of the propellant will affect the reaction rate of the chemicals inside the propellant and thus the burning rate. Between the operating range of 244 K and 344 K typical composite propellants will experience a 20 to 35% variation in chamber pressure Figure 4. Effect of oxidiser (ammonium and a 20 to 30% variation in burn time. For this reason perchlorate) particle size on burning rate in motors are left to conditions for hours to ensure the composite propellant (Sutton, Biblarz 2001, grain is uniformly at the same temperature before p504). This graph shows that a larger ignition (Sutton, Biblarz 2001, p431). Temperature proportion of smaller oxidiser particles increase effect on performance has been measured by Carter and burning rate in composite propellants. found to have a small effect on performance. In tests performed by Carter in 2008 the total impulse and average thrust of Estes rocket motors varied only by approximately 2 Ns over a temperature range of 243 K to 323 K (Carter 2008, p22). This has been noted and in order to ensure temperature does not affect results all propellants will be fired at roughly the same temperature. 2. Grain configuration Grain configuration refers to the shape of the propellant inside the motor. It can be cast into shape (case-bonded), loaded into the motor or extruded. Cartridge loaded grains are often used in small missiles and medium sized motors. They are cheaper to produce and easier to inspect. They are also easier to replace if the propellant has aged excessively. Case bonded grains are used in Figure 5. Thrust-time profiles for a number of different larger motors. They have better performance, grain configurations (ALLSTAR 2004). This figure slightly lower mass (no holding device, shows how thrust differs with time due to different grain support padding and less insulation) but are configurations. more stressed and slightly more difficult to manufacture (Sutton, Biblarz 2001, p444). Once ignited, burning precedes perpendicular to the exposed surface at a constant rate as long as the pressure remains constant (Brinley, Remington 1960, p76). Larger burning area means more propellant is being consumed in a given time which increases pressure which, as shown in Eq. (9) will increase the burning rate. Different grain configurations have different exposed area and thus will burn different as shown in Figure 5. It is important in rocket design to choose the right grain for the mission. Most hobby rockets have an end burning configuration. They have no exposed area through the core and simply burn longitudinally from one end to the other. The result of this is that thrust remains relatively constant though the burn. They are also usually cast into shape due to the lower complexity involved. This is not a problem for amateur rockets as the motor case is usually made out of inexpensive cardboard. 3. Chemical composition The main ingredients in rocket propellants are fuel, oxidiser (provides the oxygen for combustion) and in the case of composite propellants, a binder. Composite propellants contain oxidiser crystals and powdered fuel held together in a matrix of synthetic rubber or plastic. Composite propellants have been the most common class of propellants in the past three decades (Sutton, Biblarz 2001, p475). The properties of many substances were analysed to determine which ones would be appropriate to research further through experimentation. Due to the sheer number of possible propellant mixtures only a few were analysed in depth with focus being on those which have not been extensively implemented. The most popular homemade propellant consists of potassium nitrate and sucrose (sugar), known as KN/Su. Its popularity is attributed to its inherent simplicity and safety (Sugar Shot to Space). This propellant will ignite 7 Final Thesis Report 2009, ACME, UNSW@ADFA

readily, provide a fair impulse and will send a one foot rocket to a height of 1000 feet (Brinley, Remington 1960, p81). The mix will burn similar to professionally built rockets except it will have lower performance as a tradeoff to ease of manufacture (Sugar Shot to Space).Two ways it is commonly prepared is by mixing the constituents dry and packing into shape (Teleflite Corp.) or by melting the constituents and pouring the mixture into the case (Brinley, Remington 1960, p81); (Jaafar et al. 2004, p113). This mix has experimentally obtained a specific impulse (Is) of 154.8 s (Jaafar et al. 2004, p114). The simplicity to manufacture and low performance may lead people to believe there is no place for such formula but there is one such program, the Sugar Shot to Space (SS2S) which is aiming to launch a KN/Su rocket into space. One thing to be aware of is that one of the products of combustion is potassium carbonate which is hygroscopic and will form an aqueous solution of hydroxide ions; this could be quite corrosive and cause damage to the motor casing (Jaafar et al. 2004, p113). Although this propellant is quite well documented the interaction of different propellant variables has not been. By using basic chemistry knowledge a balanced equation for the reaction can be found, this is shown in Eq. (10). 5C12H22O11 + 48KNO3 > 36CO2 + 55H2O + 24N2 + 24K2CO3 (10)

Composite propellants are more closely aligned to professional applications. The shuttle rocket boosters (SRB) themselves contain a composite mix of ammonium perchlorate (AP) as the oxidiser, Hydroxylterminated polybutadiene (HTPB) as the binder and aluminium (Al) particles as fuel. While this formula is popular on professional rockets its constituent chemicals do have certain disadvantages and difficulties associated with amateur level manufacture and firing. Ammonium perchlorate is the most widely used crystalline oxidisers in solid propellants. It has good characteristics such as good compatibility with other propellant materials, good performance, quality, uniformity and availability (Sutton, Biblarz 2001, p494). Performance can be attributed to the fact that, unlike oxidisers such as potassium nitrate, it is completely convertible to gaseous reaction products (Meyer et al. 1981, p17). The main disadvantage this oxidiser has is one of the reaction products is hydrochloric acid. Each of the SRB produces over 100 tons of hydrochloric acid which is dumped directly onto the surrounding forest and ocean, significantly effecting the environment. Hydrochloric acid also serves as a nucleation site for water vapour which creates a large visible plume making the propellants use inappropriate for most military applications (Oommen, Jain 1999, p262). Ammonium perchlorate is largely inappropriate for amateur rocketry as it is a controlled substance in most countries and can only be purchased from dedicated pyrotechnic suppliers (PyroGuide). These are some motivating reasons for finding an alternative substitute. One such substitute is ammonium nitrate (AN). It has quite low performance when compared to ammonium perchlorate however it is low cost and produces a relatively nontoxic, smokeless exhaust (Sutton, Biblarz 2001, p497). The main disadvantages to its use is a phase transformation at around 32C accompanied by a volume increase which could lead to cracks in the propellant and the hygroscopic nature of the chemical (Oommen, Jain 1999, p263). Some relevant equations to the use of ammonium nitrate in propellants are shown in Table 1. This table shows the reaction pathways for the thermal decomposition of ammonium nitrate and the associated heat evolved from reaction. Reaction 4 is the reaction for detonation while reaction 5 has been suggested as the reaction which occurs on explosion (Oommen, Jain 1999, p259). By suitable choice of fuel and proper composition it has been possible to make cast compositions that yield gaseous products at a wide range of temperature, from 200C to 2000C (Oommen, Jain 1999, p262).

Table 1. Modes of thermal decomposition of ammonium nitrate (Oomen, Jain 1999, p259). Table shows the different decomposition reactions of ammonium nitrate, temperature at which they occur, heat evolved and gas volume of products.

8 Final Thesis Report 2009, ACME, UNSW@ADFA

A common binder used in professional propellants, including the space shuttle rocket boosters is hydroxylterminated polybutadiene commonly known as HTPB. It has secured its place as a popular binder due to its favourable mechanical properties. A crosslinking agent such as isophorone diisocynate (IPDI) is also added to cure the binder (Rocco et al. 2004, p804). Curing causes the binder to harden in its final form. The HTPB is also oxidised in the reaction so it acts as additional fuel (Sutton, Biblarz 2001, p501). The disadvantage of HTPB is that it is only available from dedicated rocket supply stores. Alternatives to this are binders such as Polyvinyl chloride (PVC) or silicon. PVC was used 40 years ago and unlike HTPB it does not require a curing agent to cure (Sutton, Biblarz 2001, p500). The PVC used in the experiments will be from PVC glue or PVC gel which are both essentially concentrated solutions of PVC (PyroGuide). This has been successfully used in the amateur rocket community but not in combination with ammonium nitrate. Another interesting binder is silicon. It has been reported that it works as a binder but to what extent is unknown (Nakka). The successful use of these readily available substances would be of great benefit in homemade rocket making. Finally the last major ingredient of a composite propellant is the fuel. Powdered spherical aluminium is the most popular fuel used in professional propellant. During rocket combustion the aluminium is oxidised to aluminium oxide. The aluminium increases heat of combustion, propellant density and thus specific impulse (Sutton, Biblarz 2001, p499). Aluminium also works in favor of rocket enthusiasts as it is readily available and safe to handle. The disadvantage of aluminium is that it can be quite difficult to ignite in a homemade mixture (Nakka). The difficulty comes from the tough outer aluminium oxide coating already present on aluminium due to exposure to air. Ignition will only occur once this is oxide shell is fractured or melted which can be as high as 2100 K (Beckstead 2004, p5-3). Some substances have been found to aid combustion of an ammonium nitrate and aluminium composition, these are sulfur and sodium chloride (Nakka). The alternative to this is a different metallic fuel such as magnesium. Magnesium ignites by thermal explosion rather than by a critical ignition temperature as is the case with aluminium (Beckstead 2004, p5-4). While this is used in some formulas it is quite dangerous for amateur use as it is highly flammable and explodes in the presence of moisture (magnesium MSDS). D. Measurement Each measurement system usually includes a sensing device such as a transducer, a recording device for displaying sensed information and a device for conditioning, amplifying, correcting or transforming the sensed signal into a form suitable for analysis. Recording is done on a chart or in digital form on a magnetic drive such as tape drive or disk. Range refers to the region extending from minimum to maximum rated value over which the measurement system will give a true and linear response. An additional margin is also normally provided to permit temporary overloads without damage to the instrumentation or need for recalibration (Sutton, Biblarz 2001, p721). The system which will be used to measure rocket performance in this thesis consists of a hinged cantilever beam with a load cell (Figure 6), a data acquisition system (DAQ), electronic power supply and an electronic amplifier/signal conditioner. The DAQ system consists of a software suite, channel-conditioning bus board and necessary cables. The power supply and amplifier/signal conditioner are contained in one unit. The main test rig and testing methodology was developed and used by Carter in his thesis in 2008 (Carter 2008, p1314). In order to obtain accurate results from the test rig it is necessary to have an understanding of the errors which are to be avoided throughout experimentation. There are two types of error in measurements: (1) human error of improperly reading the instrument, chart, or record and improperly interpreting or correcting these data, and (2) instrument or system errors, which usually fall into the classifications: static error, dynamic response error, drift error and hysteresis error. Static errors are usually fixed errors due to the fabrication and installation variations. This type of error can be corrected through calibration. Drift error is the change in output over a period of time, usually caused by random wander in environmental conditions. To avoid drift error the measuring system has to be calibrated at frequent intervals at standard environmental conditions against known Figure 6. Cantilever beam test rig standard reference values over a whole range. (Carter, M, p13). Instrument constructed Dynamic response errors occur when the measuring by the SACM main workshop to system fails to register the true value of the measured quantity specifications detailed by Carter while the quantity is changing, particularly when it is rapid. The rocket mounting structure and recording system should be designed to avoid excessive damping or harmonic excitation (Sutton, Biblarz 2001, p721). The cantilever 9 Final Thesis Report 2009, ACME, UNSW@ADFA

beam on the test rig was made stiff and rigid to ensure that the natural frequency of the beam was much larger than the frequency of the motors boost phase to prevent this very problem (Carter 2008, p12). The maximum frequency response refers to the maximum frequency usually measured in Hertz [Hz] at which the instrument system will measure true values. The natural frequency of the measuring system is usually above the limiting response frequency. Generally, a high frequency response requires more complex and expensive instrumentation. All parts of the instrument system must be capable of fast response. Measurements made in rocket testing are made with one of two types of instrument: those made under nearly static conditions, where only relatively gradual changes in the quantities occur, and those with fast transient conditions such as rocket starting, stopping or vibration. The later type of instrument has frequency responses above 200 Hz and sometimes as high as 20,000 Hz. These fast measurements are necessary to evaluate physical phenomena of rapid transients (Sutton, Biblarz 2001, p721). Linearity of the instrument refers to the ratio of input (usually pressure, temperature, force, etc.) to the output (usually voltage, output display change etc.) over the range of the instrument. Very often static calibration error indicates a deviation from truly linear response. A nonlinear response can cause appreciable errors in dynamic measurements. Resolution refers to the minimum change in the measured quantity that can be detected with a given instrument. Dead zone or hysteresis errors occur due to energy absorption within the instrument system or play in the instrument mechanism; in part they limit the resolution of the system. Sensitivity refers to the change in response or reading caused by special influences such as temperature and acceleration. These are usually expressed in percent change of measured value per unit of temperature or acceleration. This information can serve to correct readings to reference standard conditions (Sutton, Biblarz 2001, p722). Electrical interference or noise within an instrument system, including power supply, transmission lines, amplifiers, and recorders can affect the accuracy of recorded data, especially when low-output transducers are used (Sutton, Biblarz 2001, p722). There are several solutions to reducing noise but often a combination of methods must be used. Some simple solutions are grounding and shielding (Ott 1988, p26). Grounding is one of the primary ways of minimizing unwanted noise and in combination with correct cabling can solve a large percentage of all noise problems (Ott 1988, p73).

III.

Thesis methodology

The thesis has been separated into tasks according to appendix A which were derived from the client brief (see appendix B). The main tasks are as follows; (1) test procedures and propellants, (2) test performance of the three different compositions of propellants, (3) optimise one propellant and develop instructions for its production. The tasks were defined as such to enable easy management and progress monitoring of the project. The timeframe for these tasks are documented in the Gantt chart (see appendix C) and the milestone chart (see appendix D). Broad fuel variations which will be tested are outlined in appendix E. The cost needed for production of the propellant mixes are documented in appendix F. The composite propellant variations are based on general formulations for similar fuels given by Nakka and Orlandi which are consistent with professional rocket formulations given by Sutton (Sutton, Biblarz 2001, p19). These propellant formulations are likely to change to suit the specific substances used. The mixing process for the composite propellants will be developed during testing as it is not well documented for homemade rocketry using readily available substances. The mixing formula and technique for KN/Su rockets will be performed as in the thesis by Carter (Carter 2008, p27-30). There should be minimal change to the formulation as it has already been tested. The development of propellants is inherently dangerous, with this in mind risk assessments were conducted to cover all aspects of production (see appendix G) The static tests of the motors will be performed according to procedures outlined by Carter (Carter 2008, p13-16). Each propellant variation will be tested six times; three with the smaller C6 size case and three with the larger D11 size case. The small amount of tests is mainly due to time constraints and restrictions regarding workshop resources. The tests will be able to identify how sensitive performance is to grain size and particle size in different formulations of propellants and also compare thrust profiles of homemade rockets with Estes motors. It is expected that change in these parameters will have an effect on variables in Eq. (8) and Eq. (9) which in turn will affect the total thrust, total impulse and specific impulse.

10 Final Thesis Report 2009, ACME, UNSW@ADFA

IV.

Commercially available motors

The commercially available rockets which the homemade rockets will be compared to are the Estes brand C6 and DII black powder rockets. The sizes of the constructed cases are based on these size motors; diameters of 18 mm and 24 mm for the C and D size respectively. The thrust curves for these two motors are shown in Figure 7. Information provided by Estes shows that the C6 engine contains approximately 12.5 g of propellant while this figure is approximately 25 g for the DII engine. The C6 and DII engines have maximum impulses of 10 Ns and 20 Ns respectively; a specific impulse of 82 s in both cases (the D size contains twice as much propellant).

Figure 7. Estes C6 and DII commercially available rockets (Estes Industries 2008). Comparison between the thrust profiles for the two rockets which will be compared to the homemade mixtures. These rockets were the same tested by Carter in 2008. Carter verified that the thrust for these rockets follows the profiles given in Figure 7. Carter also tested the how temperature effects this propellants, he found that it takes a variation of approximately 30 degrees to change the burn time by less than 0.2 of a second and thrust by less than 1 or 2 N. For this reason temperature is not of concern throughout this thesis. Other information on the fuel can be estimated by knowing the propellant used is black powder. A basic black powder mix contains 75% potassium nitrate, 10% sulfur and 15% charcoal (Meyer et al. 1981, p.35). Using a program which simulates rocket propellant reactions, Propep (Lekstutis), the characteristic velocity of the propellant is estimated to be 2667 m/s while the propellant density was estimated at 1511 kg/m3. These figures will be important in further discussion relating to chamber pressure.

V.

Simulation

Two main tools were used in predicting rocket performance and that is a software package called Propep ((Lekstutis) and the isentropic relationships used in a MATLAB code (appendix H). Propep is able to find the reaction products of the propellant combustion by minimizing Gibbs free energy. Gibbs free energy is the enthalpy minus the thermodynamic temperature and entropy (International Union of Pure and Applied Chemistry). Minimising Gibbs free energy gives the reaction products of the combustion in more detail than Eq. 10. In addition to this Propep also calculates parameters such as pressure and temperature at the chamber, throat and nozzle exit as well as the specific impulse of the propellant. In order to make these calculations possible Propep uses a number of assumptions such as: One dimensional flow Zero flow velocity at nozzle inlet Complete adiabatic combustion Isentropic expansion in the nozzle Homogeneous mixing of the reactants and products Ideal-gas The thrust of the propellants was determined by the use of isentropic relationships (Anderson 2003) which resulted in Eq.11. This relationship was shown by Nakka and verified by the isentropic relationships.
2 2 1 2 +1
+1 1

= 1

2 1

+ 2 3 2

(11)

11 Final Thesis Report 2009, ACME, UNSW@ADFA

As stated earlier the thrust is optimum for p 2=p3, that is the pressure at the exit of the nozzle is equal to ambient pressure. The exit pressure (p2) is determined by the nozzle geometry and the chamber pressure (p 1) as shown by Eq. 12.
1 1

+1 1 2

2 1

+1 1

2 1

(12)

The thrust depends largely on the throat area ratio and the chamber pressure as shown by Eq.11 and Eq.12. The value of specific heats in Eq. 11 and Eq.12 is determined by the reaction that takes place and does not vary by an amount large enough to have an appreciable difference in thrust produced. Due to the manufacturing process nozzles cannot be made to the exact optimum area ratio which makes p 2=p3. For the nozzles used in the experiments plots of thrust produced with the actual area ratio and the optimum are shown in Figure 8. As the figure shows, the nozzle is optimum design, or close to optimum for a chamber pressure up to approximately 5 MPa. This pressure is larger than the expected chamber pressure for both propellants. The C and D size cases can withstand a chamber pressure of approximately 35 MPa and 27 MPa respectively. This is the result of hoop stress calculations at the thinnest and thus weakest point of the design, the nozzle thread section where thickness can be as low as one millimeter (calculations shown in appendix I).
Thrust vs Chamber pressure - C size 500 400 Optimum thrust Thrust from actual nozzle Failure of steel
Thrust force, N
500 Thrust vs Chamber pressure - D size

400

Thrust force, N

300 200 100 0

300

200 Optimum thrust Thrust from actual nozzle Failure of steel 0 10 20 30 Chamber Pressure, MPa 40

100

10 20 30 Chamber Pressure, MPa

40

Figure 8. Calculations of thrust force versus chamber pressure for C size (left) and D size cases (right). Plots based on Eq.11 and Eq.12. Failure of steel also shown and is based on hoop stress failure at the nozzle thread (thinnest section) (Callister,W 2008, p.745). Chamber pressure can be approximated with the use of Eq.13 (Nakka). This is where the difference in propellant mixtures plays a part in rocket performance. The propellant used will determine the molecular weight of products, chamber temperature, temperature coefficient, burning rate exponent, propellant density, and ratio of specific heats. The difficulty in estimating chamber pressure is finding data for the propellants temperature coefficient and burning rate exponent.
1

1 = Where

(13)

1 + 1 2

+1 1

For the propellant mix containing 65% potassium nitrate with 35% sucrose the values of temperature coefficient and burning rate exponent are 8.26 mm/s and 0.319 respectively (Nakka). Although these results do not come from a completely reliable source they do seem to be found through sound experimental logic and they are in fact the only published values. These same values will be used for the other mixtures due to lack of published data. 12 Final Thesis Report 2009, ACME, UNSW@ADFA

The characteristic velocity (c*) of the 65% potassium nitrate, 35% sucrose mixture has been quoted at approximately 950 m/s (Nakka). Propep was used to find the characteristic velocity of the same propellant for a range or chamber pressures and found it to range between 860 m/s and 920 m/s so the result of 950 m/s is quite reasonable. Differences in this result would be due to the various assumptions made in the Propep program and errors in experimental results. From data given by Propep the characteristic velocity of the aluminium based propellant is approximately 1095 m/s, it changes very little with pressure. The density of sucrose based propellant is approximately 1320 kg/m3 while the aluminium Figure 9. Plots of chamber pressure versus based propellant has a density of approximately 3 characteristic velocity for the different 1600 kg/m . With these figures in mind the plot propellants and the two size cases used. These in Figure 9 was produced. are theoretical results. It is clear from Figure 9 that both size cases using both types of propellant will result in a low chamber pressure. Taking the characteristic velocity of sucrose to be 950 m/s the chamber pressure for the C and D size cases will be 0.12 MPa and 0.15 MPa respectively. Taking the characteristic velocity of the aluminium based fuel to be 1095 m/s the chamber pressure for the C and D size cases will be 0.2 MPa and 0.24 MPa respectively. The aluminium fuel is marginally better than the sucrose fuel but this alone cannot be a measure of performance. There is also variation in chamber pressure with the two case sizes; this is due to the value of K n being higher for the D size case. It must be made clear that this is just an estimate based on the assumption that aluminium propellant has the same burning rate exponent and temperature coefficient as the sucrose propellant. The other main difference between the propellants is the specific impulse. For the C size case the sucrose based propellant has a specific impulse of 116 s compared to the aluminium based propellant, 147 s. Similarly for the D size case the impulse for the sucrose propellant is lower, 110 s compared to 138 s. These figures are once again based on calculations performed in Propep. Comparing this to commercially available rockets, Estes motors have a maximum specific impulse of 82 s (Estes Industries).

VI.

Homemade rocket construction

A. Rocket cases The rocket cases used were designed to be the same size as the commercially available Estes motors. The C size motor constructed has an outer diameter of 18 mm while the D size case has an outer diameter of 24 mm. The cases were made out of mild steel by the ACME workshop and in accordance with engineering drawings provided by Carter (replicated for convenience in appendix J).

B. Igniters Igniters need to be homemade from readily available cheap materials in order to make the homemade rockets feasible. The method used to make the igniters was a mix between that of Brinley p.154 and Sleeter p.456-461. Brinley suggested inserting a loop of nichrome resistance wire into the propellant through the nozzle; the resistance wire heats the propellant causing ignition. Sleeter suggested using some length of insulated wire, stripping the ends and passing them through a paper match head Figure 10. Final design of rocket igniter. which then go into the nozzle; the idea being that current through the bare wire causes the match heads to ignite the propellant. In both cases the igniter segment would be attached to some thicker electrical wires (lead wires) which were then connected to a battery. 13 Final Thesis Report 2009, ACME, UNSW@ADFA

Sleeters method proved quite time consuming. The wire was hard to maneuver around the match heads and stripping the wires also added to construction time. Looking at the wire another problem would be the diameter; it was simply too large and would certainly have had an effect on the nozzle p erformance. Brimleys method was problematic due to the difficulty of securing the piece of thin resistance wire into the propellant grain. Without a firm hold in the grain the igniter would be ineffective and the chances of it coming loose were too high. The final solution involved wrapping nichrome wire around a match head, a combination of both methods. This piece of nichrome wire was then attached to the main lead wires via alligator clips external to the nozzle. This worked in testing before nozzles were constructed but when used with the nozzles the wires would cross due to the small throat size and in turn short out. This problem was remedied by wrapping one side of each ignition wire in masking tape; this is shown in Figure 10. The match is left whole up until it is ready for use and the head is separated. Another advantage to this system is the wires would often break after ignition which would mean the nozzle has no obstruction during the burn. C. Aluminium fuel propellant The aluminium base propellants used a mix of 18% aluminium powder (Al), 68% ammonium nitrate (AN) and 14% binder material. The two substances used as binder material were silicone (Si) and polyurethane (PU). In order to make oxidiser suitable to use it first had to be ground with a mortar and pestle for approximately ten minutes, the comparison between the pre-ground and ground sizes is shown in Figure 11. The propellant was produced by first mixing the fuel and oxidiser substances and then adding amounts of binder material. The substances were mixed until most of the fuel and oxidiser were attached to the binder material. Although these substances are readily available they are not easy to handle. The aluminium powder was so fine that it filled the air when disturbed in the slightest. The other complication was the binder materials used; they are extremely adhesive and permanently attached to mixing implements. This would also mean lower amounts of the binder were found in the propellant due to the wastage in the mixing process. Out of the two binders used the silicone binder was the easiest to use and also resulted in finer powder than the polyurethane binder. The finer particle size means the silicone propellant would burn better due to the large surface area and more could be fit into the cases. This comparison is made in Figure 12.

Figure 11. Comparison of ammonium nitrate particles before (left) and after (right) being ground.

Figure 12. Comparison of propellant with Si binder (left) and PU binder (right)

D. Sucrose fuel propellant Sucrose fueled propellant was initially made of a mix with 60% potassium nitrate (KN) and 40% sucrose (Su) in the form of icing sugar. Further experiments tested mixes containing 55% KN and 65% KN. The stoichiometric mix, as shown by Eq.10 is 74% KN with 26% Su. An ideal mix for amateur rocketry has been quoted as 60% AN with 40% Su (Sleeter 2004, p.252) and also 65% AN with 35% Su (Nakka). The different figures are due to the reaction involving more complex products than Eq.10 provides and also the grade of substances used affecting the actual amount of chemical substance available for reaction.

14 Final Thesis Report 2009, ACME, UNSW@ADFA

The mixing method used was mixing the ingredients together as per instructions from Teleflite Corp. and Carter. This method involves measuring the required weight of fuel and oxidiser, adding in a container and then shaking the mixture for a couple of minutes. This was thought to involve less risk than melting down the reaction products to form a propellant liquid and casting into shape (Brinley, Remington 1960, p81); (Jaafar et al. 2004, p113). Another variation of the mixture was using ballmilled potassium nitrate. The ball mill consisted of four heavy spheres placed in a metal container; the container is spun at high speed to grind the powder to a smaller particle size. As discussed earlier, the smaller particle size maximises the amount of oxidiser per volume (Sutton, Biblarz 2001, p503). The ball mill was set to 400 rpm for one minute. Any longer than this and the finer powder would be clumped together in larger pieces; negating any benefit.

Figure 13. Packing of motor cases. Shows a D size motor case being packed using a wooden dowel.

E. Packing the motor cases Once the propellant was made it had to be packed into the rocket cases, this is shown in Figure 13. The propellant was placed into the two or three small spoons at a time and then damped down with the wooden dowel shown in Figure 13. Gently tapping on the end of the dowel with a hammer ensured the propellant was packed tight. The purpose of this procedure is to remove as many air voids in the propellant as possible. Air voids are areas where the local surface area is large and thus the burn rate is momentarily increased in that area according to Eq.8, this causes an unsteady burn.

VII.

Experiment method

The equipment used to measure the thrust of the rockets is shown in Figure 14. The area used was adjacent to the welding bay of the ACME building. This location was chosen as it is well out of the way of bystanders and enables a large enough safety perimeter as dictated by the risk assessment in appendix G. A detailed description of each component of the experiment follows where the item numbers are in reference to numbers in Figure 14.

Figure 14. Experiment setup outside of welding bay, ACME building. 1) Cantilever beam, 2) Amplifier, 3) Laptop/USB DAQ, 4) Digital scales, 5) Safety perimeter.

15 Final Thesis Report 2009, ACME, UNSW@ADFA

A. Cantilever beam Item 1 The cantilever beam is show in greater detail in Figure 15. The load cell used was an X-TRAN 350 N load cell. This differs to the load cell used by Carter which was an X-TRAN 250 N load cell, this difference was due to availability at the time but it was thought this would not make a large difference considering the expected forces. Each of these load cells has a resolution of 0.1%. The remainder of the rig was constructed by the ACME workshop in 2008 according to Carters design. The end of the beam is where the rocket case cup is held in place. The reason for this separation is to magnify the force at the load cell; it does this by approximately a factor of 4. Towards the end of the beam is a bolt from the base of the rig to stop excessive movement of the beam from damaging the load cell. B. Amplifier Item 2 The amplifier took input from the load cell and amplified it before it was fed to the National Instruments USB DAQ. The banana plugs in the photo provides a ground in an attempt to reduce noise (Ott 1988, p73). The voltage was set to 10 V for all experiments and this was verified with a multi meter at hour intervals in the experiment. The gain on the device was set to 200. This equipment is among the first set up as it is left to warm up for approximately 20 minutes to allow the voltage to reach a stead value C. Laptop/USB DAQ Item 3 The DAQ device by National Instruments is shown in Figure 17. This device processed the signal from the amplifier and sent it to the laptop via USB. This device has a voltage range of 10 V and a sample rate of 10kS/s (National Instruments) which is adequate for this experiment. The device was coupled with software published by National Instruments to produce plots of time versus thrust for each rocket test. The software was set to record data at a rate of 1000 Hz and then averages the data over each 10 samples. This averaging of data ensured that the plots produced were clear to read in the event of small oscillations. The software also enabled calibration of the load cell. This calibration was performed before a series of tests and involved loading the beam with 1 kg at a time and inputting the corresponding force into the software. This calibration process enabled the software to record force rather than voltage from the load cell.

Figure 15. Load cell placement under cantilever beam

Figure 16. conditioner

Amplifier

and

signal

Figure 17. NI USB-600B DAQ (National Instruments, 09). Image of the DAQ Device used.

D. Digital scales Item 4 The Digital scales used were accurate to 0.01 grams. Each rocket case was weighed empty and then labeled so they could be distinguished from one another. After this the cases were filled with propellant and weighed again. This enabled the weight of the propellant in the case to be deduced. After the propellant was tested the case was weighed again so that the weight of propellant used could be calculated. These calculations were used to compare the efficiency of the propellants.

16 Final Thesis Report 2009, ACME, UNSW@ADFA

Safety Perimeter Item 5 The safety perimeter was set at 5 m according to the risk assessment in appendix G. This safety perimeter was established at the start of the days experiments and checked before firing of the rockets. The only personnel who could enter the safety perimeter were those necessary for the experiments. E. F. Ignition system Item 6 The ignition system is shown in Figure 18. This system consisted of a 12 V, 7.0 Ah lead acid battery, some thick lead wires and some alligator clips to serve as the interface between the ignition system and the igniter described earlier and show in Figure 10.

Figure 18. Rocket ignition system. (From left to right) 12V Lead acid battery, lead wires, alligator clips, G. Experiment procedure igniter After equipment was setup according to Figure 14 the following steps were taken to collect data from the rocket test case: 1) Allow equipment to warm up for 20 minutes and verify steady voltage with a multi meter. 2) Calibrate load cell by placing weight on the cup at the end of the beam in 1 kg increments to a total of 5 kg. 3) Set the load cell overload bolt so that the bolt touches the underside of the beam when 5 kg is applied 4) Weigh the full rocket case and record the reading. 5) Place the rocket case in the end cup at the end of the beam. 6) Ensure lead wires are disconnected from the battery and the safety perimeter is clear. 7) Connect alligator clips to the igniter. 8) Zero the load cell and set the software to record data. 9) Touch the ends of the lead wires to the battery to ignite the rocket. 10) Stop recording the data. 11) Weigh the used rocket case and record the reading.

VIII.

Results

A. Aluminium fuel propellant It was unknown how safe the propellant was so small amounts were tested in a fume cupboard outside of the rocket cases. These tests were performed by placing the amount to be tested in a piece of paper, placing an igniter into the paper and wrapping the paper up to form a pouch of propellant. The success of these tests varied. On occasion the match head would ignite and simply melt the aluminium around it. Most the time combustion of the propellant would occur and the result was a bright orange flame with little visible smoke. The next phase was to test the propellant in rocket cases. Several D size rockets were tested for both aluminium based propellants and there was no successful firing. In these cases the igniter would burn and melt a case of aluminium around itself. Figure 19. Alternative aluminium To try and remedy this, the successful sucrose propellant mix based propellant preparation. was placed at the top of the propellant grain in an effort to Aluminium based propellant encourage combustion of the aluminium propellant. This was done indicated by dots, sucrose propellant in two configurations as shown by Figure 19. Neither of these by slashes. worked. Further to this the depth at which the match head of the igniter was buried into the propellant was altered with no success. Finally igniters with multiple match heads were trialed with unsuccessful results. Although successful in the lab the propellant did not combust when used in the rocket cases.

17 Final Thesis Report 2009, ACME, UNSW@ADFA

B. Sucrose fuel propellant There were a total of 22 tests of sucrose propellants, full results of these tests can be found in appendix K. These tests varied the ratio of fuel to oxidiser, the case size and the particle size of the oxidiser. For each variation in propellant there was three tests performed in a C size case and three performed in a D size case. Qualitatively all propellants produced a plume of white smoke (shown in Figure 20) and an unsteady thrust. All propellants also exhibited a sharp increase in mass flow sometimes accompanied by a load pop. This increase in mass flow was inferred through observation of the exhaust plume. The results of thrust versus time were output from the National Instruments software in an excel file. On its own these results are not useful in comparing different fuels; in each test the rocket fired at a different time in the recording and was left to run for a different time. In order to transform this data into meaningful information and find these performance parameters a MATLAB code was developed (appendix L). The purpose of the code is to take all the data collected, crop the data to only the actual burn time, integrate the plots to get impulse and take propellant mass as an input to Figure 20. Firing of calculate specific impulse. It was also able to find the burn time, maximum sucrose propellant force and average force for the burn. Typical thrust curves for each of the fuel to oxidiser ratios produce from this code are show in Figure 21. These plots are scaled from 0-1 N due to the low thrust recorded. The propellant containing 55% potassium nitrate performed poorly. The mixes containing 60% and 65% potassium nitrate produced similar thrust but the 65% mix had a longer burn time. As expected the D size cases produced more thrust, this was due to the larger chamber pressure as shown by Figure 9.
C size motors 1 65% KN 60% KN 55% KN
1 65% KN 60% KN 55% KN D size motors

0.8

0.8

Force,N

0.4

Force,N

0.6

0.6

0.4

0.2

0.2

10

15 20 Time,s

25

30

35

10

15 20 Time,s

25

30

35

Figure 21. Thrust profiles for C size rocket cases (left) and D size rocket cases (right) scaled from 0-1 N. Shows thrust for propellant formulations consisting of 65%, 60% and 55% potassium nitrate (KN). Due to experiment safety concerns the propellants which used smaller oxidiser particles were only tested in the D sized cases. The only mix tested contained 60% potassium nitrate and the results of all the trials are shown in Figure 22. The burn time for the propellant was comparable to the standard oxidiser with the exception of one trial which had a burn time of approximately 4 seconds and lead to failure of the case at the thread of the nozzle. The steady thrust for these tests were between 0.2 N and 0.3 N compared to the standard mix which had a thrust no greater than 0.2 N. The burn also seemed to be more continuous than the previous case. An interesting observation is that all propellants tested, C and D size, had a sharp increase in thrust towards the end of the burn. An example is shown in Figure 23 which shows these peaks for the D size cases. These peaks were high enough to reach the maximum force of 50 N as dictated by the load cell overload bolt. It is believed that it was one such peak which caused the failure of the nozzle in the propellant which used a smaller oxidiser particle size.

18 Final Thesis Report 2009, ACME, UNSW@ADFA

D size motors - Small oxidiser particles 1 0.8 Trial 1 Trial 2 Trial 3


30 25 20

D size motors 65% KN 60% KN 55% KN

Force,N

Force,N

0.6 0.4

15 10

0.2
5

10

15 20 Time,s

25

30

35

10

15 20 Time,s

25

30

35

Figure 22. D size rocket case containing 60% potassium nitrate and 40% sucrose.

Figure 23. Thrust peaks shown in D size tests.

The burn time was extended with a decrease in oxidiser; this is due to a higher mass of fuel in the propellant mix. The tradeoff to a longer burn time was less thrust which is clearly demonstrated in Figure 21. Other results from the tests showed similar trends to those in Figure 21.

IX.

Discussion

A. Performance of Aluminium based propellant The reason behind the poor performance of the propellant is believed to be due to the way in which the ignition system works. As stated earlier there can be problems igniting aluminium due to a harder outer case of aluminium oxide (Beckstead 2004, p.5-3). In an attempt to counter this problem a fine aluminium powder was found, the fineness of the powder was predicted to make the propellant easier to ignite as demonstrated in experiments conducted by Gurevich (Gurevich, Lapkina & Ozerov 1970). Initially it was thought that the igniter or the sucrose propellant did not burn hot enough to initiate combustion in the aluminium propellant. This theory was discarded when lab testing yielded successful combustion. The main difference between the lab and the rocket cases was the time that heat was applied to the propellant. In the rocket cases the nichrome wire heated enough to break under the weight of the alligator clips connecting it to the lead wires. In the lab tests were performed with the nichrome wires of the igniter laying horizontally, in turn they did not break as there was no force applied to them. Keeping this connection meant that current could be applied to the propellant until combustion was established and in turn this would allow the reaction to proceed. Another factor affecting the success of such propellant is the ease of manufacturing and obtaining substances. The fuel used would be difficult to find in such small particle size, it is also difficult to handle. A slight disturbance of the aluminium powder would cause a cloud of particles to spew into the air. This cloud of particles could be easily inhaled and cause discomfort and respiratory damage (Aluminium MSDS). Both the silicone and polyurethane binders used did not mix into the propellant as expected. These propellants were only expected to improve performance by a marginal value. As Figure 9 shows the propellant would only produce a slightly higher chamber pressure across a range of propellant characteristic velocities. This slight change would not result in a large different in thrust. This propellant would still be useful had it had other favourable qualities but it proved too difficult and time consuming to make in a homemade context. B. Low thrust and load cell resolution The resolution of the X-TRAN 350 N load cell used was 0.1%; 0.35 N. The cantilever beam multiplies the force on the load cell due to force at the rocket case by a factor of 4 so the resolution of the testing equipment is 0.085 N. This was considered sufficiently accurate as the force expected by the rockets was expected to be approximately 5-10 N (Estes Industries 2008). Had this been the case error for the tests would range between 0.85-1.7%. Instead the thrust produced was much lower making the error significantly higher. Further calculations were done on predicting the thrust produced by the sucrose motors. These calculations used Eq.11 to predict thrust at low chamber pressures which could be produced in the homemade rockets. The results of calculations are shown in Table 2 and show that the predicted thrust for the rockets is under 0.15 N; 19 Final Thesis Report 2009, ACME, UNSW@ADFA

only slightly bigger than the resolution of the test equipment. The results were calculated for the rocket with the ignition wires in the throat and without them as both cases were present in firing. These low values of thrust explain why the plots produced of thrust versus time show flat areas and why the plots seem to make large jumps between recorded values; the measuring equipment was not accurate for this low amount of thrust. Chamber Specific Steady state Steady Burn pressure impulse thrust state with time [s] With [s] without ignition ignition ignition wires [N] wires wires [N] [MPa] C 0.12 0.13 116 -0.36 -0.31 100+ Sucrose D 0.15 0.16 110 0.13 0.18 100+ Table 2: Predicted chamber pressure, thrust, specific impulse and burn time for sucrose based propellant, aluminium based propellant, C and D sizes and with and without ignition wires in the nozzle. The low thrust produced was due to the design of the case, nozzle and the characteristics of the propellant. The design of the case and the nozzle play a part in determining chamber pressure through the value of K n in Eq.13. For the C size case the change in chamber pressure with throat size (at constant burning area) can be seen in Figure 25. The throat used was 2.5 mm which produces a very small chamber pressure. The obvious solution is simply reducing the throat diameter to increase chamber pressure. According to Figure 8 an increase in chamber pressure would produce considerably more thrust but the reduction of the throat area must also be taken into account. In actual fact a reduction of throat size would only marginally increase performance as the narrowing of the throat would reduce mass flow through the nozzle and therefore thrust. This can be shown by some simple calculations. A reduction of the C size nozzle throat to 1 mm produces a chamber pressure of 1.8 MPa according to Eq.13. If this pressure were experienced in the C size nozzle used it would produce a thrust of 12.22 N but in order to get that pressure we have had to reduce the throat diameter to 1 mm; the effect of this is shown in Figure 24. The thrust produced would actually be much lower at 1.32 N. If the nozzle were to have the same area ratio as the original C size nozzle than the thrust produced would be 1.95 N, this is also approximately the optimum force produced at this throat size.
Chamber pressure vs throat diameter - C size
14 12

Propellant base

Case Size

Chamber pressure without ignition wires [MPa]

Thrust vs Chamber pressure - C size


60

50

Chamber pressure, Mpa

Thrust force, N

10 8 6 4 2 0 0.5

Optimum thrust Thrust from actual nozzle Failure of steel

40

30

20

10

1.5

2.5

10

15

20

25

30

35

40

Throat diameter, mm

Chamber Pressure, MPa

Figure 25. Chamber pressure versus throat diameter for C size rocket case. Based on calculations

Figure 24. Thrust versus chamber pressure for a 1mm diameter throat on a C size case. Based on calculations.

An alternative to changing the throat size is changing the burning surface area; this also increases the value of Kn and thus the chamber pressure. The advantage of this is it does not affect the mass flow rate and therefore does not experience the same drawbacks. In the example above throat diameter was reduced to 1 mm in order to obtain a chamber pressure of 1.8 MPa. This same pressure could be achieved by increasing the burning area by a factor of 6.3. There are two ways to increase the burning area; increasing the diameter of the case or changing the grain of the propellant. The internal case diameter of the C size case is 12 mm so increasing this dimension to 30 mm would increase the burning surface area by a factor of 6.3 as required. Changing the grain to a core configuration increases the burning surface area at the cost of increasing the complexity of the construction. 20 Final Thesis Report 2009, ACME, UNSW@ADFA

The final cause for the low thrust in this propellant is the characteristics of the propellant itself. The characteristics affecting the chamber pressure and thus thrust produced are the temperature coefficient, burning rate exponent, propellant density and propellant characteristic velocity. In comparing to the black powder propellant used in Estes motors, the homemade motors have a lower thrust density (1320 kg/m3 compared to 1511 kg/m3) and lower characteristic velocity (950 m/s compared to 2667 m/s). If the sucrose propellant had the same density and characteristic velocity of the black powder propellant the chamber pressure produced in the C and D size rockets would have been 0.68 MPa and 0.83 MPa (according to Eq.13). This would mean thrust would be increased to 3.8 N and 9.8 N for each of the sizes. C. Thrust peaks and nozzle failure Past experiments of sucrose based propellants also produced peaks in the thrust (Carter 2008, p.32). It was concluded that these peaks were the result of heat conduction through the metal case causing the propellant towards the bottom of the case to detonate. The problem is not an issue in commercial rockets as they tend to be disposable and made out of cardboard which is non conductive. This solution was soon discarded when further experiments in this thesis uncovered that the time at which the peaks occur is different between the same case sizes using the same propellant. If conduction was the cause it is expected that the peaks would all occur at roughly the same time which is not the case. Another explanation as to why these peaks could occur is unstable pressure oscillations during firing. These oscillations could increase in magnitude and with it rapidly increase burning rate. This phenomenon is characterized by organized oscillations occurring at well defined intervals with a pressure peak that may be maintained, may increase or may die out (Sutton, Biblarz 2001, p.348). Inspection of the thrust plots produced shows that the force spontaneously increases rather than oscillates with an increasing magnitude. For this reason unstable oscillation was ruled out as the cause of these peaks. The most likely cause of these peaks was clogging of the throat of the nozzle. As the burn progressed the ignition wires and the masking tape insulation obstructed the throat of the rocket motor and caused particles of combustion product to get lodged in the throat. The larger of these problems may in fact be the masking tape as it burns and this could cause large pieces to break off and get lodged in the throat. Another theory as to what caused the clogging of the nozzle is the angle of the converging section of the nozzle. This angle was 44 for the C size nozzle and 52 for the D size nozzle, perhaps this nozzle did not effectively channel mass flow into the throat and instead provided a junction for particles to cluster and block further flow. Clogging had the effect of reducing the throat area and thus increasing the chamber pressure as previously discussed and shown in Figure 25. For the cases that did not fail it is proposed that the chamber pressure increased due to the blockage until the blockage was freed. Once the blockage was freed the chamber pressure was at an elevated value while the throat size was back at its unblocked size which would explain why the peaks in force could cause the beam to touch the overload protection bolt configured for 50 N (shown by Figure 8). In the scenario where the nozzle failed (shown in Figure 26) the same would occur but instead of the blockage eventually coming loose it would stay lodged in the throat. Hoop stress calculations on steel with an ultimate strength of 400 MPa indicate that the C size case would fail at a chamber pressure of 35 MPa and the D size case would fail at 27 MPa. This indicates that in order for the C size case to fail the throat diameter would need to be reduced to 0.36 mm and for the D size case this figure is 0.6 mm (MATLAB code, rocket performance prediction, appendix H). It is feasible that this occurred in the D size nozzle which caused the case to fail. If the nozzle is completely blocked pressure is even higher, ranging between 413-352 MPa for the C size case and 382-401 MPa for the D size case, depending on propellant used (appendix M). Interestingly the nozzle failed as a result of longitudinal failure, had it been hoop stress a rupture in the Figure 26. Nozzle failure of D size longitudinal direction would be expected. The failure occurred case. Top half of nozzle was separated like this as a result of the thread, peaks in the thread are areas of at the thread section greater thickness compared to the valleys of the thread. With greater thickness comes greater strength which caused failure to occur along the valley of the thread.

21 Final Thesis Report 2009, ACME, UNSW@ADFA

X.

Conclusion

The thesis project has tested a number of homemade rocket propellant mixes for use in amateur rocketry. In all 22 tests were performed using test equipment developed by Carter. These tests covered variations in case size, variation in reactants and variation in fuel to oxidiser ratios. Tests found that while aluminium based propellants work well in professional application they are not suitable for homemade use. They need to have more heat applied to establish combustion and take much longer to prepare. On the other hand sucrose propellants are much easier to prepare and ignite. The thrust produced for these propellants was only a fraction of that produced by the Estes motors and the burn time was over 10 times longer. The major factors in the performance difference was the difference between characteristic velocities of the propellant and nozzle design. Another issue with the propellant was thrust peaks at the end of the burn and in one case the failure of a nozzle. These peaks and the failure were attributed to blockage of the nozzle by ignition wires and ignition wire insulation causing combustion products to clog the nozzle.

XI.

Recommendation

Future work in homemade solid rocket motors should continue to explore the design, construction and performance of solid propellants. This knowledge would benefit anyone aspiring to be involved in full scale rocketry. The research gives a general understanding of the principles behind rocket design and practical experience to highlight some of the difficulties involved. The work on homemade rocket propellants was initiated by Carter in 2009. This project represents a continuation from Carters work and follows the recommendations of further investigating homemade rocket propellant. It gives a broad understanding of homemade propellant; how to prepare, pack into cases and initiate a successful combustion. Before firing more tests it is recommended that safety concerns be further analysed and the feasibility of a more permanent test cell be researched. Future work could redesign the cases and nozzles to ensure they are safer and optimised for the sucrose propellant. From here tests could be conducted in how the propellant is prepared and packed into the case and how burn rate modifiers could be used to increase the performance of the sucrose based propellant.

Acknowledgements
I would like to thank my thesis supervisor, Dr. John Milthorpe and co-supervisor Dr. Neil Mudford for providing sound guidance throughout the thesis project. I would also like to thank the ACME workshop for construction of the rocket cases used and Mr. Andrew Roberts for his technical support in relation to electronics and data acquisition equipment used in the test equipment. Finally I would like to thank my classmates for making my time at ADFA enjoyable and my family and friends for giving me the support and motivation needed to complete my study at ADFA.

22 Final Thesis Report 2009, ACME, UNSW@ADFA

References
Anderson, J.D. 2003, Modern compressible flow: with historical perspective, McGraw-Hill. Beckstead, M. 2004, A summary of aluminum combustion, Defense Technical Information Center. Braeunig, R.A. , Basic of Space Flight: Rocket Propulsion . Available: http://www.braeunig.us/space/propuls.htm [2009, 15/04/2009] . Brinley, B.R. & Remington, B. 1960, Rocket manual for amateurs, Ballantine Books. Callister, W.D. & Rethwisch, D.G. 1997, Materials science and engineering: an introduction, John Wiley & Sons New York. Carter, M. 2008, An investigation into the combustion and performance of small solid-propellant rocket motors, Aeronautical Engineering edn, UNSW@ADFA, Canberra, ACT. Estes Industries 2008, 2008 Catalog, Estes-Cox Corp., Penrose USA. Estes Industries , Estes Educator. Available: http://www.esteseducator.com/ [2009, 14/04/2009] . Gurevich, M.A., Lapkina, K.I. & Ozerov, E.S. 1970, "Ignition limits of aluminum particles", Combustion, Explosion, and Shock Waves, vol. 6, no. 2, pp. 154-157. International Union of Pure and Applied Chemistry , Compendium of Chemical Terminology. Available: http://old.iupac.org/publications/compendium/index.html [2009, 27/09/2009] . Jaafar, M., Ali, W.K.W., Dahalan, M.N., Mamat, R., Campus, M. & Kuantan, P. 2004, "Development of solid rocket propulsion system at UTM", Jurnal Mekanikal, vol. 18, pp. 111-121. Lekstutis, A. , GUIPEP - Graphical User Interface to PEP. Available: http://lekstutis.com/Artie/PEP/ [2009, 27/09/2009] . Meyer, R., Khler, J., Homburg, A. & InterScience, W. 1981, Explosives, 6th edn, Verlag Chemie Deerfield Beach, Fla. Nakka, R. , Richard Nakka's Experimental Rocketry Site. Available: http://www.nakka-rocketry.net/propel.html [2009, 20/04/2009] . National Instruments , NI USB-6008 - 12-Bit, 10 kS/s Low-Cost Multifunction DAQ - National Instruments. Available: http://sine.ni.com/nips/cds/view/p/lang/en/nid/14604 [2009, 28/09/2009] . Oommen, C. & Jain, S. 1999, "Ammonium nitrate: a promising rocket propellant oxidizer", Journal of hazardous materials, vol. 67, no. 3, pp. 253-281. Orlandi, L. , PLASTICIZED PVC COMPOSITE PROPELLANT. Available: http://digilander.libero.it/fme/pvc.html [2009, 19/04/2009] . Ott, H.W. 1988, Noise reduction techniques in electronic systems, Wiley New York. PyroGuide , Ammonium perchlorate - PyroGuide. Available: http://www.pyroguide.com/index.php?title=Ammonium_perchlorate [2009, 20/04/2009] .

23 Final Thesis Report 2009, ACME, UNSW@ADFA

Rocco, J., Lima, J., Frutuoso, A., Iha, K., Ionashiro, M., Matos, J. & Surez-Iha, M. 2004, "TG studies of a composite solid rocket propellant based on HTPB-binder", Journal of Thermal Analysis and Calorimetry, vol. 77, no. 3, pp. 803-813. Sleeter, D. 2004, Amateur Rocket Motor Construction: A Complete Guide To The Construction Of Homemade Solid Fuel Rocket Motors, Teleflite Corp., Moreno Valley CA, USA. Sugar Shot to Space , Sugar Shot to Space. Available: http://sugarshot.org/project_description.html [2009, 19/04/2009] . Sutton, G.P. & Biblarz, O. 2001, Rocket propulsion elements, 7th edn, Wiley-Interscience. Teleflite Corp. , 5-cent sugar rockets. Available: http://balloons.space.edu/ndra/nickle.html [2009, 16/03/2009] .

24 Final Thesis Report 2009, ACME, UNSW@ADFA

Appendices
Appendix A Appendix B Appendix C Appendix D Appendix E Appendix F Appendix G Appendix H Appendix I Appendix J Appendix K Appendix L Appendix M Task breakdown structure Client brief Gantt chart Milestone chart Propellant test list Project costing Risk assessment MATLAB code, Rocket performance prediction Hoop stress calculations Engineering drawing of cases Table of results MATLAB code, Thrust analysis Complete nozzle blockage

25 Final Thesis Report 2009, ACME, UNSW@ADFA

You might also like