How Small Can a Launch Vehicle Be?

Similar documents
Analysis of Shane Telescope Aberration and After Collimation

August 3,1999. Stiffness and Strength Properties for Basic Sandwich Material Core Types UCRL-JC B. Kim, R.M. Christensen.

Modeling Laser and e-beam Generated Plasma-Plume Experiments Using LASNEX

Development of a High Intensity EBIT for Basic and Applied Science

Fission-Fusion Neutron Source

PROJECT PROGRESS REPORT (03/lfi?lfibr-~/15/1998):

Applications of Pulse Shape Analysis to HPGe Gamma-Ray Detectors

Basic Ascent Performance Analyses

Satellite Orbital Maneuvers and Transfers. Dr Ugur GUVEN

A Distributed Radiator, Heavy Ion Driven Inertial Confinement Fusion Target with Realistic, Multibeam Illumination Geometry

Flipping Bits in the James Webb Space Telescope s Cameras

Plutonium 239 Equivalency Calculations

Capabilities for Testing the Electronic Configuration in Pu

GA A23713 RECENT ECCD EXPERIMENTAL STUDIES ON DIII D

Technology of Rocket

Parametric Instabilities in Laser/Matter Interaction: From Noise Levels to Relativistic Regimes

LLNL DATA COLLECTION DURING NOAAETL COPE EXPERIMENT

HEAT TRANSFER AND THERMAL STRESS ANALYSES OF A GLASS BEAM DUMP

DML and Foil Measurements of ETA Beam Radius

Investigation of Combined Airbreathing/Rocket. Air Launch of Micro-Satellites from a Combat Aircraft

1. (a) Describe the difference between over-expanded, under-expanded and ideallyexpanded

Simulation of Double-Null Divertor Plasmas with the UEDGE Code

Optimal Control based Time Optimal Low Thrust Orbit Raising

Safety Considerations for Laser Power on Metals in Contact with High Explosives-Experimental and Calculational Results

Curvature of a Cantilever Beam Subjected to an Equi-Biaxial Bending Moment. P. Krulevitch G. C. Johnson

GA A23736 EFFECTS OF CROSS-SECTION SHAPE ON L MODE AND H MODE ENERGY TRANSPORT

RAPID GEOSYNCHRONOUS TRANSFER ORBIT ASCENT PLAN GENERATION. Daniel X. Junker (1) Phone: ,

~ _- IDOCLRXKT DMSP SATELLITE DETECTIONS OF GAMMA-RAY BURSTS. J. Terrell, NIS-2 P. Lee, NIS-2 R W.

Controlling Backstreaming Ions from X-ray Converter Targets with Time Varying Final Focusing Solenoidal Lens and Beam Energy Variation

Rocket Propulsion Basics Thrust

GA A25853 FAST ION REDISTRIBUTION AND IMPLICATIONS FOR THE HYBRID REGIME

The Gas Flow from the Gas Attenuator to the Beam Line

Partial Discharge Characterization of High-Voltage Cables and Components

Using the X-FEL to understand X-ray Thomson scattering for partially ionized plasmas

Solid Phase Microextraction Analysis of B83 SLTs and Core B Compatibility Test Units

Pathway and Kinetic Analysis on the Propyl Radical + O2 Reaction System

Astromechanics. 6. Changing Orbits

Lecture D30 - Orbit Transfers

PROBLEM SCORE Problem 1 (30 Pts) Problem 2 (30 Pts) Choose Problem #2 or #3! Problem 4 (40 Pts) TOTAL (100 Pts)

Response to Comment on "The National Ignition Facility Laser Performance Status"

CHAPTER 3 PERFORMANCE

The Launch of Gorizont 45 on the First Proton K /Breeze M

V Requirements for a Gun Assisted Launch to Circular Orbit

ION EXCHANGE SEPARATION OF PLUTONIUM AND GALLIUM (1) Resource and Inventory Requirements, (2) Waste, Emissions, and Effluent, and (3) Facility Size

An ARM SCM Intercomparison Study-Overview and Preliminary Results for Case 1

The Highest Redshift Radio Galaxy Known in the Southern Hemisphere

A NEW TARGET CONCEPT FOR PROTON ACCELERATOR DRIVEN BORON NEUTRON CAPTURE THERAPY APPLICATIONS* Brookhaven National Laboratory. P.O.

4kU. Measurement of Storage Ring Motion at the Advanced Light Source. QSTt ERNESTORLANDO LAWRENCE BERKELEYNATIONAL LABORATORY

Multistage Rockets. Chapter Notation

A New Computational Method for non-lte, the Linear Response Matrix

Study of Required Thrust Profile Determination of a Three Stages Small Launch Vehicle

GA A26474 SYNERGY IN TWO-FREQUENCY FAST WAVE CYCLOTRON HARMONIC ABSORPTION IN DIII-D

Two-Screen Method for Determining Electron Beam Energy and Deflection from Laser Wakefield Acceleration

OAK RIDGE NATIONAL LABORATORY

11.1 Survey of Spacecraft Propulsion Systems

Design And Analysis Of Thrust Chamber Of A Cryogenic Rocket Engine S. Senthilkumar 1, Dr. P. Maniiarasan 2,Christy Oomman Jacob 2, T.

Modified Kriging: Evaluation, Modification, Recommendations

Scaling of divertor heat flux profile widths in DIII-D

Chapter 7 Rocket Propulsion Physics

Gravitational Fields Review

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

INTERMOLECULAR POTENTIAL FUNCTIONS AND HIGH RESOLUTION MOLECULAR SPECTROSCOPY OF WEAKLY BOUND COMPLEXES. Final Progress Report

Principles of Rocketry

GA A22689 SLOW LINER FUSION

Impurity Transport Studies of Intrinsic MO and Injected Ge in High Temperature ECRH Heated FTU Tokamak Plasmas

PROJECT PROGRESS REPORT (06/16/1998-9/15/1998):

Plasma Response Control Using Advanced Feedback Techniques

Modeling of psec-laser-driven Ne-like and Ni-like X-ray lasers. Joseph Nilsen

Interplanetary Mission Opportunities

Scaling between K+ and proton production in nucleus-nucleus collisions *

CHAPTER 3 PERFORMANCE

GA A27806 TURBULENCE BEHAVIOR AND TRANSPORT RESPONSE APPROACHING BURNING PLASMA RELEVANT PARAMETERS

Initial Trajectory and Atmospheric Effects

Colliding Crystalline Beams

Data Comparisons Y-12 West Tower Data

CHAPTER 3 PERFORMANCE

Laser Intensity Modulation by Nonabsorbing Defects. M. D. Feit A. M. Rubenchik

12/16/95-3/15/96 PERIOD MULTI-PARAMETER ON-LINE COAL BULK ANALYSIS. 2, 1. Thermal Neutron Flux in Coal: New Coal Container Geometry

Engineering Sciences and Technology. Trip to Mars

SOLAR ROCKET PROPULSION Ground and Space Technology Demonstration. Dr. Michael Holmes, AFRL/PRSS

(4) How do you develop an optimal signal detection technique from the knowledge of

Rocket Propulsion. Combustion chamber Throat Nozzle

Magnetic Measurements of the Elliptical Multipole Wiggler Prototype

Rocket Science 102 : Energy Analysis, Available vs Required

BASAL CAMBRIAN BASELINE GEOLOGICAL CHARACTERIZATION COMPLETED

Final Examination 2015

Applied Thermodynamics - II

The temperature dependence of the C1+ propylene rate coefficient and the Arrhenius fit are shown in Figure 1.

Survey, Alignment and Beam Stability at the Advanced Light Source

Conceptual Physical Science 6 th Edition

GA A26057 DEMONSTRATION OF ITER OPERATIONAL SCENARIOS ON DIII-D

BWXT Y-12 Y-12. A BWXT/Bechtel Enterprise SMALL, PORTABLE, LIGHTWEIGHT DT NEUTRON GENERATOR FOR USE WITH NMIS

GA A23977 DETAILED MEASUREMENTS OF ECCD EFFICIENCY ON DIII D FOR COMPARISON WITH THEORY

0STI. E. Hammerberg, XNH MD SIMULATIONS OF DUSTY PLASMA CRYSTAL FORMATION: PRELIMINARY RESULTS M. J. S. Murillo, XPA

GA A22722 CENTRAL THOMSON SCATTERING UPGRADE ON DIII D

National Accelerator Laboratory

Interaction in Precipitation

GRAIL Takes A Roundabout Route To Lunar Orbit

- High Energy Gull Generator

Rocket Propulsion Overview

Transcription:

UCRL-CONF-213232 LAWRENCE LIVERMORE NATIONAL LABORATORY How Small Can a Launch Vehicle Be? John C. Whitehead July 10, 2005 41 st AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit Tucson, AZ Paper Number AIAA-2005-4506 AIAA = ASME = American Society of Mechanical Engineers SAE = Society of Automotive Engineers ASEE = American Society of Engineering Education

This document was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor the University of California nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or the University of California. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or the University of California, and shall not be used for advertising or product endorsement purposes.

How Small Can a Launch Vehicle Be? John C. Whitehead * Lawrence Livermore National Laboratory, Livermore, CA 94551 Trajectory simulations from Earth to orbit indicate comparative velocity requirements depending on vehicle size, for several propellant options. Smaller vehicles are more affected by drag, resulting in steeper trajectories that require more total velocity. Although they are technically challenging, launch vehicles smaller than 1 ton are not ruled out by the nature of ascent trajectories. I. Introduction There have been many efforts toward building ever-smaller satellites in recent years. It is therefore of interest to consider the minimum practical size for launch vehicles that can reach Earth orbit. Aerodynamic drag is more significant for smaller vehicles due to their higher ratio of drag area to mass, so trajectory analysis offers information relevant to the lower size limit. Figure 1 shows two extreme paths to orbit, each of which has two propulsive maneuvers. 1 Path A uses a horizontal launch directly into an elliptical orbit, which is circularized with a small burn. The Path B vehicle launches vertically and coasts to a stop at the desired altitude above the launch site, where it needs a circularization burn equal to orbital 60 m/s Path A 200 km not to scale Earth 7973 m/s Launch Site Fig. 1. Two extreme paths to orbit. 7790 m/s Path B 1952 m/s velocity. The labeled velocities for a 200-km orbit are based on ideal assumptions of vacuum (no drag), and impulsive velocity changes (infinite thrust). Path A ideally requires the least v, 8033 m/s, for the same reasons that a Hohmann transfer is the most efficient means of orbit raising. The inefficient Path B needs 1.21 times this v, 9742 m/s (the ratio peaks at 1.39, for 2280 km altitude). Neither extreme is actually used, but every real path to orbit lies somewhere between Path A and Path B. A trajectory can be partly characterized by the great circle arc angle and the magnitude of its circularization burn, the two of which tend to vary oppositely. Path A requires more time in the atmosphere than Path B. As the significance of aerodynamic drag increases, a preferred trajectory is pushed further from Path A, toward Path B. Higher acceleration (thrust-to-mass ratio) also favors steeper trajectories, to reduce the time spent in the atmosphere at high speeds. Low acceleration requires a steep initial ascent, to align thrust and weight vectors. Gravity losses can result in a high v requirement, even though such vehicles may pitch over and ultimately circularize at a large arc angle like Path A. Thus real trajectories are more than simple compromises between Path A and Path B. Detailed calculations are needed to estimate actual v requirements through the atmosphere with finite thrust. II. Trajectory Simulation The mathematical model was described in Ref. (1). Earth rotation and wind effects are not included. Constant thrust depletes a variable propellant quantity, followed by a coast to apogee at 200 km altitude. The launch angle is adjusted to obtain this altitude without steering. The required circularization v is added to ascent v, obtained from the propellant mass, in order to determine the total v to orbit. Starting with a set of vehicle design assumptions, the ascent propellant quantity (hence thrust duration) is varied to find the least total v. * PO Box 808, Mail Stop L-413, AIAA Senior Member. 1

Table 1 lists the relevant fixed parameters. The drag coefficient increases at transonic speeds to triple just above Mach 1, then it settles to 50% over Table 1. Parameter Fixed parameters for Mars calculations. Value the low-speed value, past Mach 4. Table 2 summarizes vehicle design assumptions for all 15 cases tried. The ratios of frontal areas to launch masses are based on propellant densities, along with the assumption of a 10:1 length-to-diameter ratio. Thrust is 1.33 of launch weight for all the liquid cases, but the solid-propelled thrust-to-mass ratio varies by only a factor of 8 for each order of magnitude in vehicle mass. Large solid rocket stages are designed to have burn areas to obtain sufficient thrust. The smallest ones typically have more thrust than necessary because the exposed Gravitational constant Earth mass Earth radius Atmospheric density at surface Atmosphere scale height (1/e decay) Speed of sound Low speed drag coefficient 6.67x10 11 N-m 2 /kg 2 5.97x10 24 kg 6380 km 1.225 kg/m 3 7.0 km 300 m/s 0.2 burn area cannot be less than the cross sectional area, i.e. an end burning grain. Table 2. Launch mass, tons Vehicle sizing assumptions for different propellants. Oxygen-hydrogen Oxygen-kerosene Solid Frontal area, m 2 Thrust, kn Frontal area, m 2 Thrust, kn Frontal area, m 2 Thrust, kn 0.01 0.1 1.0 10 100 0.021 0.10 0.46 2.15 10.0 0.13 1.3 13 130 1300 0.011 0.05 0.23 1.08 5.0 0.13 1.3 13 130 1300 0.0075 0.035 0.16 0.75 3.5 0.63 5.0 40 320 2500 III. Results Figure 2 shows the total v requirement for each simulated launch vehicle configuration. As expected, velocity varies oppositely to launch mass for all propellant combinations. Considering the simplified assumptions used in the calculations, the relative v s are more meaningful than the exact numbers. The Velocity to orbit, km/s 11 10 9 8 Oxygen-kerosene Path A, 8033 m/s Ideal escape velocity, 11189 m/s Path B, 9742 m/s Oxygen-hydrogen Solid Orbital velocity at 200 km, 7790 m/s 0.01 0.1 1 10 100 Mass at launch, tons Fig. 2. Velocity requirements to a 200-km Earth orbit. 2 graph shows that reaching orbit is somewhat more challenging for smaller vehicles. Comparisons from one propellant to another are more assumption-dependent than the effect of scaling for a given propellant option, but the general trends make sense. Hydrogen s low density increases drag to contribute to higher v requirements than for kerosene. Secondly, gravity losses are higher for hydrogen since vehicle mass falls less rapidly at higher I SP. Large solid rockets need less v because higher thrust reduces gravity losses, while scaling and density help to avoid extreme drag. However, the smallest solid rockets fare worse than the liquid options, because the scaling assumption results in excess thrust.

The above insights are supported by the magnitudes of the circularization burns in Fig. 3, and the great circle arc angles, Fig. 4. The faster-accelerating solid rockets all reach 200 km closer to the launch site, with the need for large orbit insertion burns. The large liquid vehicles Velocity to circularize orbit, km/s 10 4 10 3 10 2 Oxygen-hydrogen Path A circ. burn, 60 m/s Path B circ. burn, 7790 m/s Solid Oxygen-kerosene 0.01 0.1 1 10 100 Mass at launch, tons Fig. 3. Velocity to circularize orbits at 200 km. Great circle arc to apogee, degrees 180 150 120 90 60 Oxygenhydrogen Oxygenkerosene 30 Solid 0 0.01 0.1 1 10 100 Mass at launch, tons Fig. 4. Great circle arc angle at end point of ascent trajectories. follow trajectories that ultimately approach Path A. The extra v owed to low thrust can be called a gravity loss, since much of the propellant is consumed to react gravity rather than to accelerate the vehicle. Regardless of the thrust magnitude, the need for a nearly vertical launch, followed by almost horizontal thrusting, is inefficient for the same reason that Path B is. Figure 3 shows two apogee burns smaller than that of Path A, indicating the large hydrogen-propelled vehicles arrive at 200 km along a trajectory that has a perigee above the Earth s surface. Fig. 4 shows a saturation effect for these two cases, because 180 degrees cannot be exceeded. Hydrogen s low rate of propellant consumption results in the least average acceleration. The associated long burn times (about 300 s) permit the above-noted trajectory curvature toward Path A while still thrusting. For comparison, kerosene burn times were about 240 s, and the solid thrust duration varied from 34 s to 98 s with increasing vehicle size. The largest solid rockets have the least required v of all cases simulated, despite ascending relatively close to Path B. By way of explanation, even relatively small arc angles permit a significant horizontal velocity, e.g. 2/3 of orbital velocity for the 100-ton solid at only 12 degrees around the Earth. The distance from the launch site at 12 degrees is more than 6 times the altitude, i.e. the ascent is far from being vertical. In all the simulations having small circularization burns at large great circle positions, the total v was only weakly sensitive to changes in these two numbers. Drag data in Table 3 indicates the effect of propellant density if the two liquid propellant combinations are compared. The dramatic effect of increased acceleration on drag may be appreciated by comparing the solid propellant column to the liquid cases. The largest Table 3. Launch mass, tons Points of maximum atmospheric drag. Oxygen-hydrogen (450 s) Oxygen-kerosene (350 s) Solid (ISP = 280 s) Peak drag, kn Time, s Altitude, km Peak drag, kn Time, s Altitude, km Peak drag, kn Time, s Altitude, km 0.01 0.1 1.0 10 100 0.044 0.34 2.5 15 85 132 110 94 84 80 22.7 19.4 16.3 14.1 13.0 0.043 0.32 2.1 11 58 104 88 78 73 71 18.0 14.9 12.6 11.4 10.8 0.37 2.0 9.5 43 170 14 16 20 29 32 4.3 4.7 5.1 7.6 6.4 3

solid has only twice the peak drag of the hydrogen-propelled vehicle, while the smallest solid has almost an order of magnitude more than its liquid counterparts. The smaller solids reach their drag peaks sooner and lower than large solid rockets, owing to higher acceleration. In contrast, the liquid rockets all have the same thrust-to-mass ratios, so higher relative drag for the smaller ones reduces their actual acceleration (and velocity) enough that their drag peaks occur later and higher than for the large liquid-propelled vehicles. Note that the peak drag is one third of thrust for the smallest hydrogen case, but only one fifteenth of thrust at 100 tons. The smallest solid has the highest acceleration of all cases studied, and its circularization burn in Fig. 4 is essentially that of the ideal Path B. Its extra total velocity in Fig. 2, relative to the Path B total, is attributable to drag. Table 3 shows that the 10-kg solid vehicle s drag peak is nearly 60% of its thrust (Table 2). IV. Discussion Overall, the results underscore the fact that atmospheric drag and finite thrust lead to complexities that are not evident from Fig. 1, although that diagram is helpful for understanding the problem. Relevance of the simple non-staging model relies on the fact that most aerodynamic drag occurs relatively early, during what would be a first stage burn at liftoff thrust. The latest drag peak relative to ascent burn time occurs for the 10-kg hydrogen case, after about 40% of the total v to orbit is applied. An optimization of trajectories that includes steering, staging, and Earth rotation effects would doubtlessly yield different velocities from the simple model used here. The graphs should be interpreted mainly as approximations that indicate general trends and their rough magnitudes. Launching from an aircraft at high altitude greatly reduces the significance of drag, which in general reduces the size of the smallest possible launch vehicle for any propellant choice. The smallest solid-propelled rockets would benefit the most from a high altitude start, while liquid propellants may be preferred for the smallest ground launches. It should be cautioned that the trajectory results plotted say nothing about required propellant mass fractions or technology limits. In particular, the comparative velocities in Fig. 2 should not be interpreted as indicating the relative degree of difficulty when comparing propellant options. Considering the high ratio of area to mass on the smallest scales, non-cryogenic liquids might be appropriate. Hybrid propulsion may be of interest if the natural scaling effects offer appropriate thrust levels for desired vehicle sizes. V. Conclusion Trajectory considerations do not set a particular lower limit on launch vehicle size, but v requirements increase significantly. Achieving high propellant fractions on smaller scales is the main technical challenge, if the value of tiny payloads makes it worth doing. For example, launch vehicles smaller than 1 ton might deliver 1-kg satellites to orbit. Acknowledgment This work was performed under the auspices of the U.S. Department of Energy, by the University of California, Lawrence Livermore National Laboratory under Contract W-7405-Eng-48. References 1 Whitehead, J.C., Trajectory Analysis and Staging Trades for Smaller Mars Ascent Vehicles, J. Spacecraft, Accepted for publication, 2005. Revised from Mars Ascent Propulsion Trades with Trajectory Analysis, AIAA Paper 2004-4069, July 2004. 4