Launch Vehicle Performance Estimation:

Similar documents
The Quarter Pounder A Vehicle to Launch Quarter Pound Payloads to Low Earth Orbit

ROSE-HULMAN INSTITUTE OF TECHNOLOGY Department of Mechanical Engineering. Mini-project 3 Tennis ball launcher

Assessing Compliance with United States Government Orbital Debris Mitigation Guidelines

How to Do Flight Testing for TARC. Trip Barber NAR TARC Manager

WARM GAS PROPULSION FOR SMALL SATELLITES. J. R. French Propulsion Development Associates, Inc. ~1AlN HOUSING CATALY5THOUSP.\G C AT.

Kinematics-Projectiles

HiSST: International Conference on High-Speed Vehicle Science Technology November 2018, Moscow, Russia

In parallel with steady gains in battery energy and power density, the coming generation of uninhabited aerial vehicles (UAVs) will enjoy increased

Objective: To launch a soda bottle rocket, achieve maximum time of flight, and safely land a payload (tennis ball).

Give Wings to Imagination

THE AIRCRAFT IN FLIGHT Issue /07/12

Climbs, descents, turns, and stalls These are some of the maneuvers you'll practice, and practice, and practice By David Montoya

wind wobble unstable

Preliminary design of a high-altitude kite. A flexible membrane kite section at various wind speeds

WindProspector TM Lockheed Martin Corporation

Exploration Series. MODEL ROCKET Interactive Physics Simulation Page 01

FAI Sporting Code. UAV Class U. Section 12 Unmanned Aerial Vehicles Edition. Effective 1st January 2018

WATER ROCK. Lawndart The rocket goes straight up and comes down nose first at high speed. Disadvantages

A Simple Horizontal Velocity Model of a Rising Thermal Instability in the Atmospheric Boundary Layer

Four forces on an airplane

Table of Contents. Career Overview... 4

1. A cannon shoots a clown directly upward with a speed of 20 m/s. What height will the clown reach?

Welcome to Aerospace Engineering

Designing a Model Rocket

Honors Assignment - Vectors

Regents Exam Practice: Measurement, Kinematics, Free Fall, PJM, and UCM

QUESTION 1. Sketch graphs (on the axes below) to show: (1) the horizontal speed v x of the ball versus time, for the duration of its flight;

Rockets. Student Journal. After School STEM Academy

Aircraft Stability and Control Prof. A.K.Ghosh Department of Aerospace Engineering Indian Institute of Technology-Kanpur. Lecture-01 Introduction

Design Review Agenda

by Michael Young Human Performance Consulting

POWERED FLIGHT HOVERING FLIGHT

Advantages of Heritage Atlas Systems for Human Spaceflight

Ballistics and Trajectory

Implementing Provisions for Art. 411 of the ICR Ski Jumping

Vibration-Free Joule-Thomson Cryocoolers for Distributed Microcooling

Name: Class: Date: Multiple Choice Identify the letter of the choice that best completes the statement or answers the question.

Cutnell/Johnson Physics

First Flight Glossary

College of Engineering

TAKEOFF & LANDING IN ICING CONDITIONS

IAC-17-D The historic flight 32 of Falcon 9

UNITED KINGDOM ROCKETRY ASSOCIATION STUDY GUIDE SAFETY OFFICERS EXAMINATION AND LEVEL II FLIGHT CERTIFICATION EXAMINATION

Sea and Land Breezes METR 4433, Mesoscale Meteorology Spring 2006 (some of the material in this section came from ZMAG)

QUESTION 1. Sketch graphs (on the axes below) to show: (1) the horizontal speed v x of the ball versus time, for the duration of its flight;

ISOLATION OF NON-HYDROSTATIC REGIONS WITHIN A BASIN

DATA EQUATIONS MATH ANSWER

ME 239: Rocket Propulsion. Forces Acting on a Vehicle in an Atmosphere (Follows Section 4.2) J. M. Meyers, PhD

Anatomy of a Homer. Purpose. Required Equipment/Supplies. Optional Equipment/Supplies. Discussion

Development of Technology to Estimate the Flow Field around Ship Hull Considering Wave Making and Propeller Rotating Effects

LESSONS 1, 2, and 3 PRACTICE EXERCISES

SAFE Europe: 30 th March 2010

CHAPTER 3 TEST REVIEW

PROJECT and MASTER THESES 2016/2017

Chapter 2: Linear Motion. Chapter 3: Curvilinear Motion

Acceleration= Force OVER Mass. Design Considerations for Water-Bottle Rockets

Tilt Detection Using Accelerometer and Barometric Measurements

SULAYMANIYAH INTERNATIONAL AIRPORT MATS

Agood tennis player knows instinctively how hard to hit a ball and at what angle to get the ball over the. Ball Trajectories

SUBPART C - STRUCTURE

Exploration Series. HOT AIR BALLOON Interactive Physics Simulation Page 01

Flight Corridor. The speed-altitude band where flight sustained by aerodynamic forces is technically possible is called the flight corridor.

DEFINITIONS. Aerofoil

LAST NAME First Name(s) Student Number Practical Group as on student card as on student card Code

Exploration Series. AIRPLANE Interactive Physics Simulation Page 01

PHYSICS REVIEW SHEET 2010 MID-TERM EXAM

AP Physics 1 - Test 04 - Projectile Motion

Breaking Down the Approach

Effect of Co-Flow Jet over an Airfoil: Numerical Approach

Physics: Principles and Applications, 6e Giancoli Chapter 3 Kinematics in Two Dimensions; Vectors. Conceptual Questions

The Fly Higher Tutorial IV

IAC-06-D4.1.2 CORRELATIONS BETWEEN CEV AND PLANETARY SURFACE SYSTEMS ARCHITECTURE PLANNING Larry Bell

Rocket Activity Foam Rocket

Loads, Structures, and Mechanisms. Team C5 Matthew Marcus Chris O'Hare Alex Slafkosky Scott Wingate

Hornady 4 Degree of Freedom (4 DOF) Trajectory Program

The diagram below represents the path of a stunt car that is driven off a cliff, neglecting friction.

The exit velocity of a compressed air cannon

Unmanned Aerial Vehicle Failure Modes Algorithm Modeling

NAWGJ Education Committee

Airbus Series Vol.2 (c) Wilco Publishing feelthere.com

JAR-23 Normal, Utility, Aerobatic, and Commuter Category Aeroplanes \ Issued 11 March 1994 \ Section 1- Requirements \ Subpart C - Structure \ General

Aircraft Performance Calculations: Descent Analysis. Dr. Antonio A. Trani Professor

FAI Sporting Code. UAV Class U. Section 12 Unmanned Aerial Vehicles Edition. Effective 1st January 2019

23 RD INTERNATIONAL SYMPOSIUM ON BALLISTICS TARRAGONA, SPAIN APRIL 2007

Windmills using aerodynamic drag as propelling force; a hopeless concept. ing. A. Kragten. April 2009 KD 416

Georgian University GEOMEDI. Abstract. In this article we perform theoretical analysis of long jumps with the purpose to find

Wind Regimes 1. 1 Wind Regimes

Project 1 Those amazing Red Sox!

Practice Test: Vectors and Projectile Motion

Space Simulation MARYLAND U N I V E R S I T Y O F. Space Simulation. ENAE 483/788D - Principles of Space Systems Design

The RCM Analyst - Beyond RCM

Detailed study 3.4 Topic Test Investigations: Flight

ConcepTest PowerPoints

DESIGN AND ANALYSIS OF A COLD GAS PROPULSION SYSTEM FOR STABILIZATION

Materials: Balloon demo (optional): - balloon, string, drinking straw, flour (optional)

TWO DIMENSIONAL KINEMATICS

The Mechanics of Modern BREASTSTROKE Swimming Dr Ralph Richards

Reliability and Crew Safety Assessment for a Solid Rocket Booster/J-2S Launcher

Spring Locomotion Concepts. Roland Siegwart, Margarita Chli, Martin Rufli. ASL Autonomous Systems Lab. Autonomous Mobile Robots

Front Cover Picture Mark Rasmussen - Fotolia.com

Transcription:

Launch Vehicle Performance Estimation: John Schilling john.schilling@alumni.usc.edu (661) 718-0955 3 December 2009 Precise determination of launch vehicle performance typically requires the use of three- or sixdegree-of-freedom simulations, via established codes such as POST or OTIS. This is an oftentedious process requiring extensive technical data on the launch vehicle, payload, and mission constraints. If this data is not available, or if high precision is not available, the use of highfidelity 3- or 6-DOF simulations represents a waste of effort. In many cases, a very simple and fast method of approximately determining the performance of a launch vehicle would be preferable. Examples include: Establishing the performance trade space for a new launch vehicle Surveying existing launch vehicles for suitability to a proposed mission Determining the effects of mission changes on launcher performance Preliminary evaluation of proposed launch vehicle upgrades. Put simply, it shouldn't take more than a few minutes to estimate the performance increase associated with, e.g., a new upper-stage engine. A potentially useful technique for such analysis was developed by George E. Townsend of the Martin Marietta corporation in 1962, and published in chapter VI of the Martin-Marietta "Design Guide to Orbital Flight". Subsequent changes in launch vehicle design practice have limited the utility of this technique, but a simple refinement is identified to address the major issues. We will start with the basic question of whether a particular launch vehicle can deliver a particular payload to a particular orbit. The Townsend technique begins by assuming that all space launches consist of a direct ascent to a low circular parking orbit, followed by a series of on-orbit maneuvers to the final destination orbit. In fact, many launch vehicles fly only a direct-ascent trajectory, even to a high or noncircular orbit. However, an observation of these trajectories almost invariably finds the launch vehicle, at an altitude of a few hundred kilometers, accelerating almost horizontally through the local circular orbit velocity. One may simplify the problem by treating this as an instantaneous "parking orbit", reached by direct ascent, and with all subsequent powered flight treated as an "on-orbit maneuver". On-orbit maneuvers in the general case can represent a very complex problem, which here is greatly simplified by the fact that virtually all launch vehicle propulsion systems - even relatively low-thrust upper stages - have sufficient acceleration that an impulsive-burn approximation can be used. Determination of an optimal, impulsive two- or three-burn transfer is an elementary exercise in orbital mechanics. Launch vehicle operations may face constraints (e.g. engine restart or upper-stage coast duration limits) that prevent the use of truly optimal trajectories, but it is rarely difficult to determine (estimate?) the delta-v required for on-orbit operations.

From this, the propellant requirement for on-orbit maneuvering can be determined, and bookmarked as payload for the direct-ascent portion of the trajectory. The launch vehicle performance problem is then reduced to determining whether or not the remaining launch vehicle propellant will suffice to deliver the total direct-ascent payload (i.e. sum of actual payload, upperstage dry mass, and orbital maneuvering propellant) to the parking orbit. This is the most complex part of the problem, as the direct-ascent trajectory requires simultaneously accelerating to parking orbit velocity and climbing to parking orbit altitude while overcoming gravity, drag, engine back pressure, and steering losses. Traditionally, all of these except parking orbit velocity or specific energy are lumped together into a generic "delta-v penalty" term, typically of magnitude ~2000 m/s and strongly dependant on launch vehicle and trajectory design. Townsend determined that for most launch vehicles, this penalty term can be fairly accurately estimated as a simple function of direct-ascent flight time only. Based on a study of ~100 launch vehicle configurations in 1962, with takeoff thrust:weight ratios ranging from 1.2 to 2.0 and specific impulse from 250-430 seconds, the simple penalty estimation function is: 1. V tot = V* + 1.5E-3 T a 2 + 8.82E-2 T a + 1036 m/s where V* = sqrt(vcirc^2 + 2gH(R/(R+H))^2) T a = ascent time, s This penalty estimate is referenced to parking orbit energy; slightly more accurate results can be obtained by referencing to parking orbit circular velocity and using a penalty function of both altitude and ascent time. A fit to Townsend's graphical results gives a penalty function of: 2. V pen = K 1 + K 2 T a where K 1 = 662.1 + 1.602 H p + 1.224E-3 H p 2 K 2 = 1.7871 9.687E-4 H p H p = Parking orbit altitude, km Thus, the required performance for the direct ascent portion of the mission is reduced to a total vacuum-equivalent V of 3. V tot = (V circ + V pen - V rot ), where V rot is the contribution due to the Earth's rotation. For retrograde orbits, the latter will be a penalty. The vacuum-equivalent delta-v of the launch vehicle can be determined by applying the rocket equation to the launcher stage(s) used during this portion of the mission - remembering that a portion of the upper-stage propellant may be reserved for on-orbit maneuvers, bookmarked as payload and unavailable for use here. Vacuum performance figures should be used for all engines, even those used primarily at low altitude. Parallel staging (e.g. strap-on boosters) will complicate this somewhat, but the calculation is still relatively simple. Concurrent with this calculation, one must calculate the time required for the direct-ascent phase of the flight. This will for the most part be a simple burn time calculation, though for best results

staging and ignition delays should be considered. If a vehicle configuration allows for an optional coast before upper-stage ignition, the most accurate estimate generally comes from using the minimum coast period. With the ascent time and parking orbit altitude known, equation 1 gives the total estimated penalty delta-v. If the parking orbit altitude is not known, a value of 100 nm or 185 km will serve as a reasonably accurate estimate for most modern launch vehicles. If the penalty delta-v satisfies equation 3, the launcher has sufficient performance for the proposed mission. If necessary, the calculation may be repeated with a new payload estimate, iterating until closure. The Townsend approximation for the penalty term assumes that all of the various penalty effects can be reduced to a simple function of the ascent time, which seems deceptively simple. However, most of the "penalty" relates to events occurring early in the flight, e.g. drag and gravity losses. The chief constraint on a vehicle designer's effort to overcome these losses, is the time spent in this regime. If the vehicle flies a relatively uniform acceleration profile, total ascent time will correlate strongly with the time spent in the near-vertical, high-drag regime where most loss effects are found. One further source of 'penalty' delta-v, is trajectory shaping losses. These are most often due to a requirement that the early trajectory loft the vehicle to a sufficiently high altitude, sometimes higher than the parking orbit altitude, that it will not fall back into the atmosphere before reaching orbital velocity. This is even more strongly correlated with total ascent time; a longer ascent period requires greater excess vertical velocity. However, Townsend's work explicitly assumed multistage vehicles with identical stage mass ratios, T:W ratios, and specific impulses. Given that assumption, the correlation between total direct-ascent time and early flight behavior will be quite good. The resulting launch vehicle, will not be so good, and will not reflect modern design practice. If dissimilar staging is allowed, the optimal design is generally one which uses a high-thrust first stage (often augmented by strap-on boosters) to rapidly escape the early high-loss flight regime, and a relatively low thrust but high Isp upper stage to most efficiently add delta-v once most loss terms have become irrelevant. Thus, for example, an Ariane 5G launch vehicle requires an astonishingly long 1,500 seconds to fly a direct ascent LEO mission, in spite of a very high 1.9:1 initial thrust:weight ratio. Clearly, for such a vehicle the total ascent time is not a good proxy for early flight behavior. The Ariane 5, and many other modern launch vehicles, accelerates briskly through the high-loss regime, then almost leisurely works its way up to orbital velocity. And its behavior is very, very poorly predicted by this model in its basic form. A simple refinement, however, restores the utility of the model. Instead of using the absolute time required to accelerate to local circular orbit velocity, we can use the weighted average of A: the true acceleration time, and B: the acceleration time of a hypothetical three-stage-to-orbit vehicle where each stage has the same mass ratio, and the vacuum I sp and initial T:W ratio of the actual vehicle at launch. Such a vehicle will behave quite similar to the actual vehicle in the critical early period of the trajectory, but will more accurately match the underlying assumptions of the Townsend Model. The equivalent ascent time of the hypothetical three-stage vehicle is, 4. T 3s = 3 [ 1 e (-0.333 Vp / g Isp) ] g I sp /A 0

where V p = delta-v to parking orbit, m/s A 0 = acceleration at launch, neglecting gravity and atmospheric effects, m/s 2 Based on a survey of over 1,000 data points representing the known performance of 17 modern launch vehicles to a variety of destination orbits, the best-fit modification of equation 2 becomes: where, 5. V pen = K 3 + K 4 T mix where K 3 = 429.9 + 1.602 H p + 1.224E-3 H p 2 K 4 = 2.328 9.687E-4 H p H p = Parking orbit altitude, km 6. T mix = 0.405 T a + 0.595 T 3s This model predicts the effective penalty delta-v with an RMS error of ~260 m/s. With the total delta-v of a space launch mission ranging from ~8,500 m/s (ideal LEO launch) to 13,000 m/s or more (Earth escape or GEO payload delivery), this gives a <3% error in total mission delta-v and usually <10% error in payload capacity. It is possible to do better still, for launchers whose performance is known in at least some cases. Comparing known performance with that predicted by the model, one often sees a systematic offset that can then be reasonably applied to the same launcher in a new configuration or application. In some cases, offsets can be identified for specific components of a modular launcher. For example, an examination of 210 data points reflecting the performance of the Delta IV launch vehicle family, most of them from the Boeing Delta IV payload planner's guide, shows generally good agreement with the model. The accuracy is greatly increased, with an RMS error of only 113 m/s in mission delta-v, if we add an offset of +314.8 m/s to the core vehicle, -70.5 m/s to each GEM-60 solid rocket booster, and 119.4 m/s to the optional 5-meter payload fairing. This suggests that the Delta IV family was designed for optimal performance with the full compliment of boosters, and is somewhat less efficient in with just the core vehicle. Also, unsurprisingly, the larger payload fairing has a non-trivial performance penalty due to increased drag. With this additional knowledge, it is possible to very accurately predict the performance of the Delta IV to orbits not listed in the payload planner's guide, or to examine the impact of potential changes to the vehicle configuration (e.g. a twin-engine upper stage) Finally, it is worth noting some of the limitations of this technique. It requires only basic engineering data on the rocket stages in use, but that data must be accurate - the survey of current launchers conducted during this work, revealed several cases where published engineering and/or performance data was clearly inaccurate; often because bare motor weights were quoted for upper stages that in fact incorporate substantial additional hardware. It is also necessary to accurately model the orbital maneuvers conducted in the final stage of a payload launch. These have not been discussed here, and for the most part represent a simple problem in orbital mechanics - especially for launches to circular orbits in a plane directly accessible from the launch site, where a near-impulsive Hohmann transfer can be performed. As

previously mentioned, however, there may be operational constraints on these orbital maneuvers which must be taken into account. It is possible to "cheat" and predict unrealistically high performance, by positing strap-on boosters or even entire first stages which provide high thrust for an extremely short period. This would extrapolate to a very short burn time per equation 6, but the expected performance gains will not materialize if the vehicle reverts to a low-thrust trajectory too early in the flight. Great skepticism should be exercised regarding any vehicle whose first stage and/or SRBs burn for significantly less than one minute. A related issue concerns fast-burn boosters, especially those derived from surplus ICBMs. The original Townsend model shows net delta-v losses reducing monotonically with ascent time, down to a period of two hundred seconds. The more complete survey here confirms this, and suggests the extrapolation might continue down to perhaps 100-150 seconds. But a sufficiently high-acceleration launcher will reach impractically high velocities while still in the lower atmosphere, resulting in excessive drag losses. This is not considered in the present model, and is a potential source of error. And, obviously, this is a technique which in its present form is limited to ground-launched, ballistic, rocket vehicles. An extension to air-launched vehicles is possible, and may be implemented in the future. But vehicles which use air-breathing propulsion, or which make significant use of aerodynamic lift during the ascent phase, will use substantially different trajectories and suffer very different penalties - at present, there seems no alternative but a comprehensive 3DOF trajectory simulation for such vehicles.