https://ntrs.nasa.gov/search.jsp?R=19940020279 2020-07-10T14:57:50+00:00Z NASA Contractor Report 4568 Approximate Optimal Guidance for the Advanced Launch System T. S. Feeley The University Los Angeles, and J. L. Speyer of California California at Los Angeles Prepared for Langley Research Center under Grant NAG1-1090 National Aeronautics and Space Administration Office of Management Scientific and Technical Information Program 1993 Abstract A real-time guidance scheme for the problem of maximizing the pay- load into orbit subject to the equations of motion for a rocket over a spheri- cal, nonrotating Earth is presented. An approximate optimal launch guidance law is developed based upon an asymptotic expansion of the Hamilton-Jacobi- Bellman or dynamic programming equation. The expansion is performed in terms of a small parameter, which is used to separate tile dynamics of the problem into primary and perturbation dynamics. For the zeroth-order prob- lem the small parameter is set to zero and a closed-form solution to the zeroth- order expansion term of the Hamilton-Jacobi-Bellman equation is obtained. Higher-order terms of the expansion include the effects of the neglected pertur- bation dynamics. These higher-order terms are determined from the solution of first-order linear partial differential equations requiring only the evaluation of quadratures. This technique is preferred as a real-time on-line guidance scheme to alternative numerical iterative optimization schemes because of the unreliable convergence properties of these iterative guidance schemes and be- cause the quadratures needed for the approximate optimal guidance law can be performed rapidly and by parallel processing. Even if the approximate solu- tion is not nearly optimal, when using this technique the zeroth-order solution iii PI_A_OtNi; P_G[ 8(.ANK NOT FH.14ED always provides a path which satisfies the terminal constraints. Results for two-degree-of-[reedom simulations arc presented for the simplified problem o[ flight in the equatorial plane and compared to the guidance scheme generated by the shooting method which is an iterative second-order technique. iv Table of Contents Abstract iii Table of Contents V List of Tables viii List of Figures ix List of Symbols xi 1. Introduction 1 o The Peturbed Hamilton-Jacobi-Bellman Equation 5 2.1 Expansion of the H-J-B Equation ................. 8 2.2 Solution by the Method of Characteristics ............ 10 2.3 Determination of the Optimal Control .............. 11 2.4 Determination of the Forcing Functions .............. 12 1 Modelling of the ALS Configuration 14 3.1 Equations of Motion for the Launch Problem ........... 16 3.2 Propulsion .............................. 18 3.3 Aerodynamics ............................ 18 3.4 Mass Characteristics ........................ 21 3.5 Gravitational and Atmospheric Models .............. 22 V 3.6 Expansion Dynamics ........................ 24 3.6.1 Two-Dimensional Flight .................. 25 0 Zeroth-Order Optimization Problem 27 4.1 Optimization Problem Statement ................. 27 4.2 Zeroth-Order Coordinate Transformation ............. 29 4.3 Zeroth-Order Analytic Solution in the Cartesian Frame ..... 31 4.4 Linking the First and Second Stage Subarcs ........... 36 o First-Order Corrections 40 5.1 Correction to the Lag-range Multipliers .............. 41 5.2 The First-Order Forcing Function ................. 41 5.3 Relating the Partial Derivatives of the Wind Axis Frame to the Partial Derivatives of the Cartesian Frame ............ 43 5.4 Partial Derivatives of the Analytic Solution ............ 44 5.4.1 Partial Derivatives of Some Common Terms ....... 44 5.4.2 5.4.3 Partial Derivatives of the Analytic States ......... Solution to the Linear System of Unknown Partials 45 . . 48 , Aerodynamic Effect along the Zeroth-Order Trajectory 52 6.1 Inclusion of an Aerodynamic Effect in the Zeroth-Ordcr Problem 53 6.1.1 Zeroth-Order Aerodynamic Effect in the Rectangular Co- ordinate System ....................... 56 6.1.2 First-Order Correction Terms ............... 59 Results for the Rectangular Pulse Punctions ........... 60 Aero Pulses in the Body-Axes Frame ............... 62 vi 7. Results 67 o The Relationship equation between Calculus of Variations and the HJB 83 8.1 Correction Terms to the Lagrange Multipliers .......... 83 8.2 Expansion of the Euler-Lagrange Equations ........... 87 8.2.1 Expansion of the State Equations ............. 88 8.2.2 Expansion of the Lagrange Multiplier Equations ..... 89 8.3 Expansion of the Boundary Conditions .............. 91 8.3.1 Expansion of the Transversality Conditions ........ 92 8.4 Solution to the First-Order Problem ................ 93 8.5 Solutions to First-Order Linear Partial Differential Equations.. 95 8.6 Formulation of First-Order Correction Terms for the ALS Probleml00 8.7 Results ................................ 105 9. Conclusions 114 A. Zeroth-Order Solution for Three-Dimensional A.1 Zeroth-0rder Coordinate Transformation Flight ............. B. Canonical Transformations C. Point Inequality Constraints D. Analytic Partial Derivatives BIBLIOGRAPHY for Zeroth-Order Solution 117 124 129 133 137 142 vii List of Tables 3.1 Vehicle Mass Characteristics .................... 22 7.1 Comparison of Results ....................... 72 7.2 Comparison of computation time ................. 81 8.1 Comparison o[ open loop results .................. 106 8.2 Comparison of closed loop results ................. 106 °o, VIII List of Figures 3.1 ALS Vehicle Configuration ..................... 15 3.2 Coordinate Axis Definition ..................... 17 3.3 First Stage Drag Model ....................... 19 3.4 First Stage Lift Model ....................... 19 3.5 Second Stage Aerodynamic Model ................. 21 4.1 Transformation of Coordinal_e Systems .............. 30 6.1 Coordinate frames for the aerodynamic pulse functions 6.2 Model for aerodynamic pulses in x-direction ........... 6.3 Model for aerodynamic pulses in z-direction ........... 6.4 Open loop zeroth-order path for body-axes aerodynamic ..... 55 57 57 pulses . 66 7.1 Hamiltonian versus Angle-of-Attack first stage .............................. at continuous points of the 69 7.2 First stage model for the drag coefficient ............. 70 7.3 Comparison of the first stage and second stage aero models along the vacuum path .......................... 71 7.4 Angle-0f-Attack vs. Time ..................... 74 7.5 Thrust Pitch Angle vs. Time ................... 75 7.6 Altitude vs. Time .......................... 76 7.7 Velocity vs. Time .......................... 76 7.8 Flight Path Angle vs. Time .................... 77 ix 7.9 Dynamic Pressure vs. Time .................... 78 7.10 Velocity Lagrange Multiplier vs. Time .............. 79 7.11 Flight Path Lag-range Multiplier vs. Time ............ 80 8.1 Geometric Interpretation of Integral Surface ........... 98 8.2 Open loop solution for Lagrange multipliers at staging conditions 108 8.3 Open loop solution for Lagrange multipliers at first stage initial conditions .............................. 109 8.4 Closed loop solution for flight path angle Lagrange multipliers 110 8.5 Closed loop solution for velocity Lagrange multipliers ...... 111 8.6 Closed loop solution for angle-of-attack .............. 112 List of Symbols English Symbols a, b, c CD CD_ C Dc.2 CDa3 CL CL_ C L_,2 cq C_,,Cw Cw D f(y,_,T) f, f_ constants of the quadratic mass equation drag coefficient linear coefficient in the drag model quadratic coefficient in the drag model cubic coefficient lift coefficient in the drag model linear coemcient in the lift model quadratic coefficient in tile lift model side force coefficient constant terms associated with the Lagrange for the velocity components u, w multipliers constant term used to rewrite the Lagrange multipliers in terms of mass, C_, = _--_,rmo+ C_ second stage value of Cw given first stage initial conditions drag force primary dynamics the i th term of the asymptotic expansion of the primary dynamics partial derivative of the primary dynamics with respect to the control u xi g g_ C(y, u, t) h hi hf.p_c he H H Opt [f w_nd HLH HI H_, Isp J K(Q,P,t) L perturbation or sccondary dynamics the i th term of the asymptotic expansion of the perturbation dynamics partial derivative of the perturbation with respect to the control u dynamics gravity sea-level gravity scalar function of the augmented altitude final attained altitude performance index specified final altitude atmospheric density scale height the Hamiltonian of the systcm the optimal Hamiltonian the Hamiltonian of the wind axis system the Hamiltonian of the local horizon or Cartesian the Hamiltonian evaluated at the final time system first derivative of the Hamiltonian with respect to the control u second derivative of the Hamiltonian with respect to the control u specific impulse performance index Hamiltonian for a new set, of variables Q and P lift force xii L_L rnf 17_s_ge t Yns_ge2 M N(y,t) p P P(x,t) P= P, Ptt Q Lagrangians used in Appendix B mass of the vehicle final mass specified mass at end of first stage before staging specified mass at beginning Mach number; M = rE_ 303 number of engines of second stage after staging dynamic pressure equality constraint appears in Appendix C the partial of the dynamic pressure equality constraint generalizcd coordinate of old system in Appendix B generalized coordinate of new system in Appendix B the optimal return function starting at the initial conditions the partial derivative of the optimal return function with respect to the initial state x the partial derivative of the optimal return function with respect to the initial time t i th term of the asymptotic expansion of the primary dynamics the partial derivative of the i Lh term of the expansion of the optimal return function with respect to the initial state x the partial derivativc of the i Lh term of the expansion of the optimal return function with respect to the initial time t dynamic pressure generalized coordinate of old system in Appendix B side force in Chapter 3 on ALS modelling generalized coordinate of new system in Appendix B .o0 XIU T Te 80S S S(q,Q,t) t, to tl _s_ge T rl T_ T_ U, V vl /'f a pec X X (x,Y,Z) Y radial position of the vehicle: re q- h radius of the Earth the forcing function associated with the i _a correction term speed of sound Cross-sectional area of the combined vehicle generating function defined in Appendix B initial time final time stage time total thrust of the vehicle value of the thrust for the first stage value of the thrust for the second stage vacuum thrust per engine the i th term of the asymptotic expansion series of the control velocity components velocity associated with the inertial frame final attained velocity specified final velocity initial states downrange Position coordinates state vector for the right-handed inertial frame xiv Greek Symbols /3 X 6(c, h) A Amst,,ge "/f. pec angle-of-attack; control in the wind axis system vehicle sideslip angle; control in the wind axis system velocity heading angle ratio of the atmospheric density to the small parameter discriminant associated A = 4ac- b2 with the quadratic mass equation discontinuity in the mass at staging the small expansion parameter; ratio of the atmospheric scale height to the radius of the Earth the jth power of the small expansion parameter flight path angle final attained flight path angle specified final flight path angle Lagrange Lagrange multiplier multipliers associated associated with the state y with the wind axis states Lag'range multipliers Ah, Ax, Ay, A._ associated with the Cartesian states # velocity roll angle; control in the wind axis system Lagrange multiplier associated with the terminal constraint ft(y(tst_9e)) ¢ constraint latitude imposed by the staging condition of the rocket ¢2(q,p,t) ¢(yf, Ts) new generating function equal to S(q, Q, t) scalar component of performance index on y Xv _(_) P p_ p_ O" T 0 vector of terminal constraints atmospheric density sea-level atmospheric density reference atmospheric density specific fuel consumption time longitude pitch angle; control in the Cartesian system Miscellaneous nm sin C08 tan sinh -i _(m) a( ) _() _() _dT--_() _() _o) ( Symbols nautical mile sine function cosine function tangent function inverse hypcrbolic sine function argument of the inverse hyperbolic sine function the differential of ( ) the time-varying variation of ( ) the variation of ( ) with time held fixed denotes the time derivative of ( ) with respect to the independent variable time partial derivative of ( ) with respect to the independent variable mass partial derivative of ( ) with respect to the initial state x xvi _ot( ) ), 0 )s )o ), lira partial derivaLive of" ( ) with respect to the initial time t prime superscript used for second stage values which are linked to the initial conditions on the first stage subarc subscript denotes the initial conditon of ( ) subscript denotes the final conditon of ( ) superscript denotes the optimal ( ) subscript denotes sea-level value; subscript denotes the characteristic direction in Chapter 2 limit operation xvii Chapter 1 Introduction An approach to real-time optimal launch guidance is suggested here based upon an expansion of the Hamilton-Jacobi-Bellman or dynamic pro- _amming equation. In the past, singular perturbation theory has been used in expansion techniques used to solve optimization problems [1, 2, 3]. For singular perturbation methods the states are split up into a set of 'fast' and 'slow' variables. The solution is then sought in two separate regions; one re- gion where the fast states are dominant and an outer region where the slow states are determined. A composite solution can then be determined by com- bining the two solutions. Matching asymptotic expansions is one method for obtaining the final solution. This research uses a regular asymptotic expansion which is assumed valid over the entire trajectory of the launch optimization problem. An example of a launch optimal control problem is to determine the angle-of-attack profile which maximizes the payload into orbit subject to the dynamic constraints of a point mass model over a rotating spherical Earth. The solution of this type of optimization problem is obtained by an iterative optimization technique. Since the convergence rate of iterative techniques is difficult to quantify and convergence is difficult to prove, these schemes are not suggested to be used as the basis for an on-line real-time guidance law. In contrast, an approximation approach is developed which is based 2 upon the physicsof the problem. Thrust and gravity are assumedto be the dominant forcesencounteredby the rocket while the angle-of-attackis usually kept small in order to minimize the effect of the aerodynamic forces acting on the vehicle. Numerical optimization studies [4] havebeen performed which support this assumption. These results also indicate that ignoring the aerodynamic pitching moment has a negligible effect on the performanceof the vehicle. Thus the launch problem would seemto lend itself to the useof perturbation theory. It is shownthat the forcesin the equationsof motion can be written as the sum of the dominant forces and the perturbation forceswhich are multiplied by a small parameter c, where ¢ is the ratio of the atmospheric scale height to the radius of the Earth. The motivation for this decomposition is that for ¢ = 0, the problem of maximizing the payload into orbit subject to the dynamics of a rocket in a vacuum over a fiat Earth, is an integrable opti- mal control problem. The perturbation forcing terms in the dynamics producc a nonintegrable optimal control problem. However, since these perturbation forces enter in with a small parameter, an expansion technique is suggested based upon the Hamilton-Jacobi-Bellman equation. The expansion is made about the zeroth-order solution determined when c = 0. This zeroth-order problem is now solved routinely in the generalized guidance law for the Space Shuttle [5] with a predictor/corrcctor along the desired path. scheme employed to guide the vehicle The higher-order terms of the expansion are determined from the solution of first-order linear partial differential equations which require only integrations which are quadratures. Quadratures are integrals in which the in- tegrand is only a function of the independent variable. Previous so]ution meth- 3 ods applied to guidance problems have motivated the approach suggested here. These include the explicit gnlidance laws, E-galidance, developed by George Cherry [6] for the Apollo flight. By writing the dynamics strictly as functions of the independent variable a solution was obtained by quadrature integra- tions. Past applications [7, 8] of the proposed scheme, have shown that very close agreement with the numerical optimal path is obtained by including only the first-order term. Because no iterative technique is required, this scheme is suggested as a guidance law since the quadratures can be performed rapidly. Chapter 2 contains a general formulation of the perturbation prob- lem associated with the Hamilton-Jacobi-Bellman partial differential equation (HJB-PDE). The technique for determining the higher-order expansion terms due to the perturbation forces caused by the atmosphere Earth model is discussed. Lastly, the recursive relationship and the spherical for the control is presented. In Chapter 3, the characteristics for the Advanced Launch System (aka National Launch System) and the general equations of motion in terms of the small parameter e, are given. For e = 0, a simplified optimal launch problem in the equatorial plane is formulated, and its solution in terms of elementary functions is given in Chapter 4. The coordinate system transformation used to obtain the analytic solution is included. Also discussed is the linking of the trajectory subarc for the first stage to the subarc of the second stage. In Chap- ter ,5 the first-order correction term to the control is determined. Results are presented in Chapter 6 and compared to the shooting method solution, which is a numerical iterative second-order optimization technique. It was found that during much of the first stage the aerodynamics are not small when flying the optimal vacuum trajectory. Chapter 7 presents a method for reshaping the zeroth-order trajectory by including an aerodynamic effect. This effort centers on the useof constant aerodynamicpulse functions which are obtained by averaging the aerodynamicsalong the zeroth-order path during various time intervals. Lastly, Chapter 8 relates perturbation theory and the Calculus of Variations with the expansionof the Hamilton-Jacobi-Bellman equation. Tile equivalenceof the two solution methods is presented. The Peturbed Chapter 2 Hamilton-Jacobi-Bellman Equation The optimal control problem can be formulated as one which mini- mizes a performance terminal constraints; Minimize index subject to a set of nonlinear that is, dynamics and a set of J= (2.:) with the dynamics = f(y, u, r) + _9(y, u, r) (2.2) subject to the terminal constraints qJ(yf, Tf) ---- 0 (2.3) and the initial conditions y(t) = x = given (2.4) Note that Y is an n-dimensional state vector, u is an m-dimensional control vector, _ is a small parameter, r is the independent variable, _) =a dy/d'r, t is the initial value of the independent variable, and x is the initial state at t. namics. Eq. (2.2) is separated into two portions: primary and secondary dyNote that the control appears in both parts. The primary dynamics 5 can be assumed to dominate over the secondary dynamics because the secondary dynamics are multiplied by the small parameter (e) and therefore have a small perturbing effect on the system. The Hamilton-Jacobi-Bellman (H-J-B) equation [9] is - Pt = H °pt = min H = p_[/o_t + cgOpt] (2.5) uEbt where/4 is the class of piecewise continuous bounded controls and u_t(x, P_., t) is obtained from the optimality condition H_ = 0 and from the assumption that the Legendre-Clebsch condition is satisfied (H_,_, is positive definite). In addition, fopt =_ f(x, uOpL, t) and gore _ g(x, uOpt,t). The Hamilton-Jacobi- Bellman equation will be used to determine minimizes the cost criterion J. the optimal control policy which The function P(x, t) is called the optimal return function and is de- fined as the optimal value of the performance index for a path starting at x and t while satisfying the state equations (2.2) and the terminal constraints, i.e., P(x,t) = ¢(yl,r/) at the hypersurface Bellman partial differentional equation _P(y/,'r/) = 0. The Hamilton-Jacobi- (2.5) can be interpretated [10] as the derivative of the optimal return function P. The optimal return function is a constant since it is dependent only on the terminal conditions and thus the total derivative of the optimal return function along an extremal path must be zero. dP Pt + p_[fovt + cgOpt] 0 dt Each point in space belonging to the optimal trajectory must give the same value to the optimal return function as the optimal P(x, t) since the trajectory is considered optimal from thc initial conditions (x, t) to the terminal manifold. Now, if a non-optimal control is chosen at any point in the trajectory, then the resulting terminal state, as generated by' the system equations, must produce a value for the optimal return function equal to or greater than the optimal value. Thus the control that minimizes the cost is the control which at each point of the trajectory causes the derivative of the optimal return function to be zero. This is the fundamental notion represented by the Hamilton-Jacobi-Bellman equation. Note that x and t can be either the initial or the current state and time, respectively. In this context, it will be used to represent the current state and time. Also note that ew._ry admissible constraints qJ(Yl, rl) = O. trajectory must satisfy the terminal P(z, t) can be expanded ,as a series expansion in e as ,_'(_,t)= _ f',(_, t)_' i=O (2.6) and the optimal control can also be expanded in a series expansion as oo _°_(_, &,t)= _ _,(_,t)_' i=0 (2.7) where u _t is obtained by substituting Eq. (2.6) into Eq. (2.7) and expanding the function. Therefore, it is possible to obtain the control law in feedback form. The zeroth-order control, Uo, is the optimal control for the zeroth- order problem where e = 0. If an analytic solution can be obtained for the zeroth-order problem then higher-order solutions for the control can be ob- tained by expanding the Hamilton-Jacobi-Bellman equation P, = Z P,,(_, 0 _'= - i----O F,_(_,t)_' f,_' + _g,_' i=O i= 1 (9.8) 8 where the dynamics have been expressed as expansions of the form OC f°Pt(m, u °m, t) = _ f_(x, u, t)d i=0 f"(x, t)= i=O (2.9) (2.1o) Expanding Eq. (2.8) and collecting terms of equal powers in e, produces the following set of linear, first-order, partial differential equations Pit + P_zf_ t= = i-I -_ j=o Pjz(fi-j _- gi-j-l) R4(z,t, ei-l,...,Po) i= i,2,... (2. ii) The expansion next section. of the Hamilton-Jacobi-Bellman equation will be detailed in the 2.1 Expansion of the H-J-B Equation The solution to the optimal control problem requires the evaluation of the Lag-range multiplicr, P_. Note that the quantity P_ is the partial derivative of the optimal return function with respect to the state y at the initial time or the current time (since at r = t, y = x). The function P= is expanded in a series in the small paramcter e. The terms of this series expansion, P_=, are evaluated in terms of quadrature integrals which are functions of P_. Recall that the functions P_ require the previously evaluated terms Pj=, f,_j, and g__j_ l for j = 1,...,i - 1. The coefficients f, and gi are the i it' term in the series expansion of f and g given in Eqs. (2.9)-(2.10). Since f and g are assumed to be sufficiently differentiable, they are expressible in a power series in e in terms 9 of the conLrol. For a scalar control, this yields g°Pt(x, It °pt , t) = 0U i x,t,_=0 _ uje 3 (2.13) The above equations assume that the zeroth-order control, uo, is the dominant term in the series (Eq. (2.7)). This implies that the higher-order correction terms, 7zl, _z2, ..-, have a much smaller ef[cct on the optimal return flmction, [_(x, l), than the zeroth-order term. rFhe first ['our terms of f and g are obtained by use of [']qs. (2.12) _n(i (2.13). fo -- f°m(x, Tzo,t)= f(x,_zo, t) fl = &- f3 -- utf_(x, uo, t) zt 2 _f_(x, uo, t) +u2f_(x, uo,t) tt 3 -j f_,_,_(x, Zto, t) + zt,Tz2f_,(:c, Uo, t) +u_f_(_, _o,t) (2.14) (2.15) (2.16) (2.17) 9o = 9°_(_, _,o,t) = 9(x, _o,t) gl = ulg,,(:c,uo,t) _ - 2 _""(_:'_o,t) + _9,,(x, uo,t) g3 - 6 g,,_,(X, Uo, t) + Ulu2guu(X, +u39,,(x, uo, t) Uo, t) (2.1s) (2.19) (2.20) (2.21) lO Note that in taking the partials with respect to u in Eqs. (2.12) and (2.13), the partial is taken first and then the partial is evaluated at x, t with c set equal to zero. In other words, the partials arc evaluated along the zeroth-order path. 2.2 Solution by the Method of Characteristics The H-J-B equation (Eq. (2.5)) is a first-order partial differential equation. The expansion of the H-J-B equation results in the first-order differential equation for P_ stated in Eq. (2.11) with the boundary condition P_(xl,tl) = 0, for i = 1,.... Recall that f_t denotes the dynamics of the zeroth-order problem (e = 0) using the zeroth-order control u = u0. Recall also that the forcing term /_ is only a function of expansion terms of P of order less than i. The method of charactcristics is used to solve a set of linear or quasi- linear partial differential equations. cation and solution of characteristics This technique [11] requires the identifi- curves. The characteristic direction ds is defined by the equation Pi,(dT)s + P_,(dy), = (dP_)., i= 1,o,, ..- (2.22) Eqs. (2.11) along with (2.22) can be put in the form (ayL = (aP, L The characteristic directions for Eq. (2.23) are given by the solution of the differential equation that is obt'ained by setting the determinant of the matrix given in Eq. (2.23) equal to zero, such that (dy)s- fo(d'r)s = 0 ==_ (dy/dv), = fo (2.24) 11 The subscript s denotes tile characteristic direction. Therefore, the charac- teristic curves of the equations, zeroth-order optimal trajectory for any order term of P/, are given by the 90 = f0 (2.25) whose solution is denoted as yo(r; x, t). The solution for P/ is given by P,(x, t) = - fit, R°dT where /_ is defined along the zeroth-order path as (2.26) R °= l_(yo,r, Pi__(yo,r),',Po(Yo, r)), i= 1,2,... (2.27) Thercfore, having already dctermincd P terms of order less than i, a solution for P, can be determined by integrating R4 from the current 'time' to the final 'time' along the zeroth-ordcr path. 2.3 Determination of the Optimal Control Since the primary and secondary dynamics, f and g, are expanded in terms of the control (Eqs. (2.12) and (2.13)), the control expansion terms u0, ul, u2, ..-, need to bc determined. The optimality condition provides the necessary tool to obtain these control tcrms. It can be stated as By expanding Px[f_ + eg_] = P,= ei (fi_ + eg,.)e' = 0 -- i=0 (2.28) and multiplying out the terms of the two power series and equat- ing like powers of e, the following relations are obtained e° : P0. £ = 0 (2.29) 12 12 +&.[9,, + u:f..] + P2.f. =0 (2.30) (2.31) Note that uo, the optimal control for the zeroth-order problem, can be solved using Eq. (2.29). Similarly, ul can be solved using Eq. (2.30) and u2 can be solved using Eq. (2.31). 2.4 Determination of the Forcing Functions Eqs. (2.14)-(2.21) and (2.29)-(2.31) can be used to solve for the forcing functions Ha where Eq. (2.11) can be restated as i--I Ha= - Z PJ_(f,-J + ._t,-,-,) j=O i = 1,2,... Using the above equations, RI is (2.32) R, = - &.(f, + o0) = -&.(u,L + g) (2.33) With the use of the optimality condition of Eq. (2.29), R_ becomes & = - &=go (2.34) Similarly, the equation for It2 is R2 = -- Po.(f2 + gl) - Pl=(fl + go) R2 simplifies to the following equation when Eqs. (2.14)-(2.21) are substituted into the previous equation. u_ D R2 = --_, o=L_, - Pl_go (2.35) and (2.29)-(2.30) (2.36) Finally, R3 can be expressed as R3 = -Po.(f3 +g2) - P,.(f2 +gl) - P2.(fL +go) This simplifies to 13 (2.37) = ,,r:'.:o,+-ULU 2 _ 1 U 1 ,-,.U[l go+(2y.3I8).] Using the expression for Ri, the expression ers, Pi., can be expressed as for the Lagrange multipli- - OOPx, - fits O-_Pz_ dr + _lt_-Ot _1_, OOtxI (2.39) Once these P,, are determincd, they can be used in the optimal control ex- pansion (Eq. (2.7)). As made apparcnt in the above equations, the solution becomes increasingly complex as thc higher-order correction terms rely on the state information from the lower-order trajcctories. Modelling Chapter 3 of the ALS Configuration This chapter presents the modelling characteristics and the equations of motion for the rocket. Included are sections on the properties of the propul- sion, aerodynamics, masses, gravity, and the atmosphere. A small expansion parameter, the ratio of the atmospheric scale hc'ight to the radius of the Earth, is then used to separate the dynamics into the primary and perturbation ef- fects. Lastly, the equations of motion for the zeroth-order a vacuum over a flat Earth are presented. problem of flight in The Advanced Launch System (ALS) is designed to be an all-weather, unmanned, two-stage launch vehicle for placing medium payloads into a low Earth orbit. The spacecraft (fig. 3.1) consists of a liquid rocket booster with seven engines and a core vehicle that contains three engines. All ten liquid hydrogen/liquid oxygen low cost engines are ignited at launch. Staging occurs when the booster's seven engines have exhausted their propellant. The three core engines burn continuously from launch until they are shut down at or- bital insertion. Launched in the equatorial plane and ending at the perigee of a 80nm by 150nm transfer orbit, the flight occurs in two-dimensions over a nonrotating, spherical Earth. Note, the booster is assumed to ride on top of the core throughout the first stage trajectory. 14 15 3315.2 Liquid RocketBooster 2667.2 Core Vehicle 1737.2----- _l 1497.2._......_ _ "_ 1683.2 1516.6 I 4 50.9 0.0 .... ii Stations Measured From Exit Plane in Inches Figure 3.1: ALS Vehicle Configuration 16 3.1 Equations of Motion for the Launch Problem The general equations of motion for a launch vehicle modelled as a point mass over a spherical, nonrotating Earth are given for flight in threedimensions as h Vsin7 = (T cos_ cos_ - D) - g sin y m = [- (T cos a sin/3 - Q) sin # + (T sin a + L) cos/z] mV V g +[(To+hi _]cos7 [(Tcos_sinB-Q)cosl_ + (T sin c_ + L) sin/_] = (mV cos_) V tan ¢ cos y cos X 4 (re+h) = (Vreco+s h"f)ccoos sX¢ _) = rh = V cos ")"sin X (re +h) -aT.,c (3.1) (3.2) (3.3) (3.4) (3.5) (3.6) (3.7) The vehicle coordinate system is shown in figure 3.2. Note, the engines are not gimbaled and the aerodynamic pitching moments are neglected. For a vertical launch Eqs. (3.3)-(3.4) experience a singularity caused by the velocity being zero and by a flight path angle of 90 degrees, respectively. Therefore, a pitch- over maneuver must be made at launch and equations different coordinate frame must be used. of motion written in a 17 L V Y D mg Figure 3.2: Coordinate Axis Definition 18 3.2 Propulsion Thrust is assumed to act along the centerline of the booster-core vehicle configuration and to be the same constant value for each engine. The total thrust of the rocket changes after staging as the seven engines of the booster are discarded, leaving only the three engines of the core vehicle. T = (T,_c - npA_) T,,_ = n x 580, 110. lbs. where T,,,c is the total value of the thrust when acting in a vacuum and the number of engines is n = l0 for the first stage and n - 3 for the second stage. Notice the variation of the thrust due to the atmospheric pressure p is given for an undcrcxpanded nozzle and thus a conservative value for thrust is used. The value of the engine nozzle exit area is A_ = 5814.8/144. sq ft. The specific fuel consumption of the rocket is =l sea I_p g_ ft (3.8) and the specific impulse I_p = 430. seconds. after staging occurs. The value of a remains the same 3.3 Aerodynamics Since sideslip causes drag, the vehicle is assumed to fly at zero sideslip angle, so that only the angle-of-attack gives the orientation of the vehicle rel- ative to the free stream. The direction of the lift vector is then controlled through the velocity roll angle. With no sideslip, the side force Q is identically zero. Therefore, 19 :_ 0 Alpha o_---_ 0 8 0 Mach Figure 3.3: First Stage Drag Model I0 \ Alpha Mach Figure 3.4: First Stage Lift Model 20 L = Ct.qS, D = Ct)qS, Q = CQqS = O (3.9) where CL, Co, CO. are the lift, drag, and side force coefficients, respectively, S is the cross-sectional 1 2 area of the combined vehicle (booster + core), and q = ipV is the dynamic pressure. The cross-sectional area S is assumed to be the same constant value before and after staging occurs. The aerodynamic data has been provided in tabular form [4] and is modelled by polynomials in a with Mach-number-dependent coefficients. For the first stage, the aerodynamic coefficients arc written as CD(M, ol) = Coo(M) + CD 2(M)ol 2 + CD 3(M)c_ 3 CL(M,o_) = CL_(M)c_ (3.10) where the Mach-number-dependent terms have been obtained from cubic-spline curve fits of the tabular data. Three-dimensional plots [12] of the first stage drag and lift models are shown in Figmres 3.3 and 3.4. Note that the drag coefficient of this vehicle at supersonic and hypersonic speeds has a minimum at a positive angle of attack as shown in Figure 3.3. This is caused by the aerodynamic shielding of the booster by the flow field of the core. After staging, the vehicle operates in the hypersonic the aerodynamic force coefficients are modelled as flow regime and CD(OI) ----- CDo Jr- CD,_ Ol -t- CDc, 2Ot 2 CL(a) = CL.a + CL _a 2 (3.11) with constant coefficients CDo = .2011, CD,_ = 0.0, CD,_2 = .001811, CL_. = 21 0.4 I w Ct_ 0.35 0.3 0.25 0.2 -10 -5 ! 0 ct (deg) ! ,I 0.5 .... Ct, , 0.25 -0.25 -0.5 5 10 Figure 3.5: Second Stage Aerodynamic Model .039962, and CL2 = .00100272. vided in figure 3.5. Tile aerodynamic plot of CL and CD is pro- 3.4 Mass Characteristics The inert weights of the booster and core, the weight of the propellant, the payload and payload margin, and the weight of the payload fairing comprise the ALS takeoff weight. The fairing encases the payload and is carried along by the core vehicle until orbital insertion. The vehicle mass and sea-level weight characteristics are shown in Table 3.1. The time at which staging is to occur is obtained from the first stage mass flow rate and the propellant of the booster rr_-o_tt,_,,t = 153.54 sec. tstage _- 7aT,_c 22 Vehicle Stage Vehicle Component Core Booster Core + Booster Inert Mass Propellant Payload Payload Margin Payload Faring Total Core Inert Mass Propellanl: Total Booster Total at Take-off Take-off Weight (lbs.) 176,130.00 1,479,180.00 120,000.00 12,000.00 39,120.00 1,826,430.00 216,880.00 1,449,980.00 1,666,860.00 3,493,290.00 Table 3.1: Vehicle Mass Characteristics where the vacuum thrust per engine is T_o_ = 580110. Once the stage time, tile total first stage mass flow rate, the takeoff weight, and the inert weight of the booster are known, then the weight of the vehicle at the end of tile first stage and the initial weight in the second stage can be calculated. For this vehicle the values are msao,1 = 1421890. lbs., mst_oc2 = 1250010. lbs., Amst_gc = 216880. lbs. 3.5 Gravitational and Atmospheric Models The gravitational acceleration is modelled as an altitude-varying tion by the inverse square law, r2 e g = g"(re + h)2 func- 23 but will be assumed constant in the zeroth-order problem to facilitate obtaining an analytic solution. The constant values for gravity at sea-level and for tile radius of the Earth are ft g_ = 32.174 -- see 2 re = 2.09256725 x 10 r ft. The atmospheric density is expressed by the exponential function, p = pre-(r¢+h)/ho = pre-rJh, e-h/h, = pse-h/ho (3.12) where he is the atmospheric scale height and ps is the sea-level reference density. The values for these parameters are p, = .002377 slugs h., = 23,800. ft. ft 3 The form of the density is chosen to motivate the selection of a small parameter to exclude chosen as the aerodynamics in the zeroth-order dynamics. If e is e = hs/rc (3.13) and defining then by atmospheric isfies the requirement small, i.e., _5(e,h) = p(e,h) e (3.14) properties ¢5(e, h) > 0. Tile exponential density also sat- [3] that the perturbation term in the dynamics remains lim 6(e, h) --+0 _---+0 (3.15) Satisfaction of this property used in the launch problem. will allow more general atmospheric models to be 24 The atmospheric pressure is "also expressed as an exponential function, p -- p_e -h/% (3.16) where hp is the atmospheric pressure scale height and p_ is the sea-level reference pressure. The values for these parameters are lbs Ps = 2116.24 f-_ hp = 23,200. ft. The speed of sound can be obtained by thc relationship SOS _ W_ with the specific heat ratio for air given as F = 1.4 . The gravity can be rewritten as g=g_- gsh(2r_ + h) (r_+h) 2 =gs- egsh(2r_ + h)r_ hs(r_ + h) 2 (3.17) where the expansion parameter has formally been introduced and the second term is clearly small in comparison to the first term which is the value for gravity at sea-level, g_. 3.6 Expansion Dynamics In terms of the small parameter are rewritten as c, the full-order equations of motion V sin 7 (3.18) cos c_cos/3 - 9_ sin 7 m npA.r_ g_h(2r_ + h)r_ sin 3' +_ cos a cos f_ + mh, hs(r, + h) 2 P SV2CDre 2mhs (3.19) ] ] 25 T"V (cos o_ sin L¢sin # - sin ol cos #) .QsCOS-7 V rzp Ae re -- e m--m--_.' (cos a sin ,g sin # - sin a cos #) + pSVr_.,,., e--t_Q 2ruh, sin # + Ct, cos #) +e -re+h + .qs V(r_+h) cos _] g Tvac _ (cos c_ sin _ cos # + sin oesin #) mV cos 7 npA_r_ -emvh, cos I' (cos a sin/3 cos # + sin c_ sin #) (3.20) J r pSVr_. Vr_. tan 0 cos "7cos X] +e Lm-_z,T;-.y7(ocs,. sin _ - CQcos ;_) + h,(re + h) (3.21) V croes-CTOcoSs0 _(1 - _/_) Vc°s-TsinX(lre - e/@.) (3.22) (3.23) Where the binomial formula has been used to rewrite (r_+h)-l and latitude since re >> h. for the longitude 3.6.1 Two-Dimensional Flight In this section the three-dimensional equations of motion are reduced for flight in a great-circle plane (the X-Z plane) over a flat, nonrotating Earth. If the vehicle is assumed to be restricted to fly in the equatorial plane then the lift, thrust, and velocity vectors all lie in the same plane and the roll angle (# = 0) is eliminated from the equations. Under the previously mentioned assumptions of no side force (Q = 0) and no sideslip (_ = 0), the zeroth-order equations of motion representing flight in a vacuum over a flat Earth become h = Vsin'7 (3.24) 26 9 -- TVQC cos a - g_ sin 7 m _ Tt, ac gs mV sin a - _- cos 7 V cos 7 - re rh = -aT,_ _ m = mo - aT,,,,c(7- - To) X = Xo = 0.0 ¢ = ¢0 = 0.0 (3.25) (3.26) (3.27) (3.28) These are the system dynamics used to obtain an analytic solution zeroth-order optimization problem presented in the next chapter. to the Zeroth-Order Chapter 4 Optimization Problem The solution to the zeroth-order a coordinate transformation. A canonical optimization transformation problem is derived by from the wind axis to the rectangular or local horizon coordinate frame allows the zeroth-order problem to be solved analytically. The solution is in closed form up to some constants that can be determincd numerically to solve the two-point boundary value problem. The conditions for connecting the second stage subarc to the first stage subarc are then prcsented. 4.1 Optimization Problem Statement In this section the zeroth-order optimization The problem is to maximize the payload into orbit problem is presented. J = -rrt$ subject to terminal constraints on the altitude, velocity, and flight path angle, h/ = hl,,,,o , Vf = Vfop,_, "),I ="tlo_, subject to the state discontinuity in the mass at a interior point where staging Occurs j 7_stage2 _ ?T_staqel -- /_sta9 e 27 28 and subject to the equations of motion for flight in tile equatorial plane. h. = Vsin7 = --cosa-9_sin'_ - O- £n = T 9_ mvSina-_c°s'7 V cos 7 re -aT _ rrl = trio -- aT(T -- TO) (4.1) (4.2) (4.3) (4.4) (4.5) Note, in this section and when discussing the zeroth-order trajectory, the total vacuum thrust will be represented by T and the subscript notation will be dropped. The Hamiltonian for this system can then be expressed as H= AhVsinT+Av(Tcosa-g.,sinT)+ m T 9s A-r(_---_ sin a - K cosT) (4.6) The zeroth-order control law determined by the optimality conditon is T H_, = -TAmr sina + m-V'%cosa = 0 (4.7) By the strengthened Legendre-Clebesch condition H_,_ > 0 choose x, tanol -- VAv COS _ ---- VAv + sin a = X'r (4.8) + Whereas the optimal control can be derived in terms of the states and Lagrange multipliers, an analytic solution is not possible for the states and Lagrange 29 multipliers written in the wind axis frame. Therefore, a coordinate transformation into the Cartesian reference frame is presented in the next section. In section 4.3 an analytic solution is obtained using this transformation. 4.2 Zeroth-Order Coordinate Transformation The analytic solution for the zeroth-order problem can be found in the Cartesian coordinate system but the equations of motion of the full sys- tem which include the aerodynamic forces are written in the wind axis system. Therefore, to derive the zeroth-order control and the first-order correction to the control the transformation of coordinates and especially the transformation of the Lagrange multipliers must be known. This can be accomplished by a canonical transformation [see appendix B] from the (0, ¢, h) coordinates to the right-handed coordinate system (X, ]i, Z), where X is positive in an eastward direction along the equator, Z is positive pointing is orthogonal to the X - Z plane. The relationship towards the Earth, and Y between the two reference frames (see figure 4.2) is X = re0, Y = re¢, and Z = -h. In two-dimensions, the corresponding velocity coordinates (u,w) are considered positive in the pos- itive X and Z directions, respectively. A necessary and sufficient condition [13] for a canonical transformation is the equivalence of the Hamiltonians in the two reference frames. HLH = AxdX + AvdY + Ahdh + A,_du + A_,dw Hw_,_ = AodO + A,d¢ + Ahdh + AvdV + A._d'y (4.9) (4.10) 30 ___[__!d_ I w _./....f T Body Axis Axis Local Horizon =X Inertial Reference Frame Figure 4.1: Transformation of Coordinate Systems 31 This equivalence is obtained through the Jacobian of the transformation. fore, the transformation u = V cosT, w = -Vsin7 There- (4.11) requires and thus, This produces A._ - V sin 7 - V cos 7 ]Aw the transformation of tile Lagrange multipliers, Av = A_cosT-Awsin7 A-r = -V(A,,sinT+A_cosT) Ao = T_Ax Ae = reAy and the transformation of the states, V = v/u 2 +w 2 1// sin7 - V (4.12) (4.13) (4.14) (4.15) (4.16) (4.17) 4.3 Zeroth-Order Frame Analytic Solution in the Cartesian In this section an analytic solution will be derived for the zeroth-order problem of maximum payload into orbit for flight in a vacuum over a fiat Earth. This solution is made possible by the coordinate transformation presented in 32 the previous section. The equations are of motion in a Cartesian coordinate fraxne .]( _- u ? = o_Y=Yo=O h = -_ T = -- cOS0p 7?2 iJ = _b - O_v=vo=O T sin 0p + g_ Tt2 rh = -aT ==_ m =mo - aT(T -- TO) (4.18) (4.19) (4.20) (4.21) (4.22) The Hamiltonian is H = Axu - AhW + A,_T cos0p + A,_( -T sin0p + 9_) m m The zeroth-order control law is determined by the optimality conditon (4.23) Hop - T A,, sin 0p - T)% cos 0p = 0 m m (4.24) Therefore, comes using the strengthened Legendrc-Clebesch condition the control be- tan0p - COS _p sin0p = A,, A_ + A,, + (4.25) 33 The Lagrange multipliers are obtained using J_y £x = 0 i_ = 0 ;(. = -,Xx _" = X_ with the boundary conditions where _x, r'h, _., v_ are unknown Lagrange multipliers associated with the ter- minal constraints. For the unconstrained downrange problem, the solutions to the adjoint differential equations are -_x = tl X = 0 Ah = r'h A_ = v,,=C,, A_ = C,_ + kh(T-- T0) (4.26) (4.27) (4.28) The equations of motion can be integrated by changing the independent vari- able from time to mass and using the mass equation (Eq. (4.5)) to substitute mass for 7-. As a consequence, the Lagrange multipliers are rewritten as ._,_ = C,, m _ + _ = c__+ _ + _ (4.29) (4.30) (4.a) 34 where c- (aT7 b- 2 AhC_, aT a = C +VL -- mo c,,, = C,,,+ Ah-j-_ (4.32) (4.33) (4.34) (4.35) The derivatives of the states with respect to mass are du C,, - dm amx/cTr_ 2 +bm + a dw _ A,_ 9s dm am_/cTn 2 +bm + a aT dX u - _ dm aT dh w - -- dm aT (4.36) (4.37) (4.38) (4.39) Note that c > 0, a > 0, and the discriminant of the quadratic A _=4ac-b 2>0since 4 A- (aT) 2 (AhC_) 2 mass equation (4.40) From these differential equations the solution is found from standard integrals. u = Uo av/'a sinh -l \ m_v/_ ] - sinh -l \ too v/._ ]J = ,_o- _T(m- too) (4.41) aC%-/-'_a [sinh-' {<2_ma%_/__+b_n]]--sinh-I {k2a_+-bnmoo-]-]_ ]J h gs (m- too)_ + (m- too) ho 2(aT)2 aT Wo (4.42) 35 ma(af_,/- E sinh-l \ m_/_ _ ] _ sinh- , \ 7-r__v_ C_' [sinh-' (2crn + b) - sinh-' (2_/_+ (_- _o) X No -- //,0 aT C,, \ / \ 7 o /j [sinh-'( 2a+bm] - sinh-' (2a+bm0'_] (4.43) a(crT)v/-C \_ sinh-I The equation for the altitude common terms. can be manipulated further to eliminate some h ho 2(efT) 2 (m - too) -k WO aT -mG(_r)2v_ sinh-1 a(#-_Vv_/-,d [Lsinh-' (\2_arav4/--_brn'_) sinh-I (\2a+7-bnmoov_ )] G(_AT, )2c [_/Cm2o+bmo+a_x/cm2+bm+a ] At the final time, H I = -1 by tile transversality condition. Using the tlamil- tonian and the three state equations u,w, and h, which have prescribed initial and final values, the four unknown constants associated with the two-point boundary value problem can be solved. For the problem of flight restricted to a plane, the unknowns are mj,, C_,, C_,, and Ah. The analytic state equations (Eq. (4.41)-(4.43)) are nonlinear and thus no statement can be made about the 36 existence solutions or uniqueness of the set of constants found. Therefore, if multiple are found tile solution set which minimizes the Itamiltonian would be chosen. At the very least, the Legendre-Clebesch weak relative minimum must be satisfied. condition, H,,,, _> 0, for a 4.4 Linking the First and Second Stage Subarcs Of interest in this section is the linking of the two subarcs of the two-stage rocket. By the corner conditions, the Lagrange multipliers for all the states must be continuous. (4.44) The analytic solution previously presented is still valid for either subarc but only by using this relationship between the Lagrangc multipliers can the sec- ond stage be connected to the first stage subarc. Recall that the constant C_, is associated with the initial condition of the Lagq'ange multiplier for the ver- tical velocity component. For a subarc with first stage initial conditions, the equations become A_,(t) = ),,o(t,t_g_)+ ,\a(t - t._t_g_) t _>t_t_v_+ (4.46) Rewriting the Lagrange multipliers using the corner condition replacing time as the independent variable, results in and with mass Ah = r,h = constant to < t < tf A,, = u,, = C,, = constant to O, c' > 0 and the discriminant because A'=4a'c'-b a=4\aT2] C_>0. (4.56) 38 The simplified form of the solution to the state equations (Eqs. (4.41)-(4.43)) is also still valid but with the first stage subarc used as the initial conditions of the second stage subarc. u uo-_---_ sinh-_ C_, Is L _sinh -_ {2a' + b'ms_g_2_ 1 (4._7) 0,11 /2a' + b'rnstaqe2 -_2T2x/'d k_---_)- sin},- k-- _ = _ m2o) 9,(m _ -- = TrZsta_e2 ) h ho + 2(_T,)2 2(_T_) 2 T/%W m0"UJ0 (4.58) e_ [sinh-' /2a+brn.,t,_v_,'_ {2a+bm°_l Ah x/Urn = + b'm + a' - _/d _t,_'2 + [tm,L_,v_2 + a' -_ a(_T=)2d - rLstagc i • +-_(o:q)_c 39 These are the equations that result from linking the first stage subarc to the second stage subarc. These equations will be used to evaluate the states at a time after staging occurs when the initial time is before staging. The first-order correction terms will require the analytic solution for the states at any future time along the zeroth-order trajectory. Chapter 5 First-Order Corrections The use of the asymptotic expansion of the dynamic programming equation as discussed in Chapter 2 by the approximate optimal guidance scheme is an improvement over past analytic techniques whose guidance laws were lim- ited to operate in tile exoatmospheric region [6, 14]. The higher-order correc- tion terms of the HJB expansion can bc used to compensate for tile effects of the atmospheric forces neglected in tile exoatmospheric mination of the first-order correction to tile zeroth-order solution. The detercontrol is the subject of this chapter. As noted before, tile solution to the first-order optimization problem requires only the integration of quadratures, which can be evaluated quickly enough to permit this method to be implemented as a real-time guid- ance scheme. The correction to the I,agrange multipliers and thus tile cor- rection to the control is constructed in the following sections. Also derived are all the partial derivatives needed to evaluate the quadratures. The partial derivative chain rule is employed since the analytic solution is found in the Cartesian frame while the first-order forcing function, Rl, used to evaluate tile quadratures is expressed in the wind axis frame. Recall that the angle-of-attack is the control variable and tile aerodynarnic coefficients are modelled as func- tions of the angle-of-attack. For this reason tile perturbation dynamics are left expressed in the wind axes frame. 4o 41 5.1 Correction to the Lagrange Multipliers The higher-order in Eq. (2.26). terms of the optimal return fimction were presented P, (:r, t) = - _tl ROdr By taking the partial derivative of this integTal the correction Lagrange multiplier can be caleulatcd. Recall, term to the & 05 _ - Oz ft " OR 5000 -- shooting method .... zeroth-order - - -first-order L /_ _ T I I I 25 75 125 175 225 275 325 375 time Figure 7.7: Velocity vs. Time 77 6O I J J ] I I 5O \ t_0 _9 :_,\ 4O '_ \',\ , _) 3O -- shooting method ..... zeroth-order - - - first-order ..... first-order w/pulse 2O 10 0.0 25 I I I I I I 75 125 175 225 275 325 375 time Figure 7.8: Flight Path Angle vs. Time 78 1500.0 I 1000.0 v ¢9 oO ¢D ¢2, _9 °_,,_ 500.0 I I 1 I 1 I --shooting method ..... zeroth-order - - - first-order ..... first-order w/ pulse \ N_\ ' 0.0 I I I I I I 25 50 75 100 125 150 175 200 time Figure 7.9: Dynamic Pressure vs. Time 79 0.0 I f I I I I -0.5 "7 t.) -1.0 hO J ,! I -1.5 I > I I J -2.0 I I J shooting method ..... zeroth-order - - - first-order ..... first-order w/pulse -2.5 25 i I I i I I 75 125 175 225 275 325 375 time Figure 7.10: Velocity Lag-range Multiplier vs. Time 80 3000 2000 1000 I I I I I I J •e • "/_ 9t J-, ¢9 hD -1000 -2000 -3000 -4000 25 shooting method ..... zeroth-order - - - first-order ..... first-order w/pulse I I J I I I 75 125 175 225 275 325 375 time Figure 7.11: Flight Path Lag'range Multiplier vs. Time 81 Method CPU time (see) zeroth order 49. first vacuum 304. first pulse shooting 344. 426. Table 7.2: Comparison of computation time agreement with the optimal solution. A last point about these result is that the inclusion of the rotation of the Earth in the problem is expected to continue to reduce the time of flight and consequently orbital insertion. increase the final weight available at The convergence of the asymptotic expansion is indicated by the re- sult of the first-order solution in comparison with the shooting method so- lution, thereby precluding the need to include higher-order correction terms. This convergence is tentative since it took the inclusion of the aerodynamic pulse functions in the zeroth-order problem to achieve the best results. Alas the convergence properties when using these pulses cannot be guaranteed or even quantified. Finally, since this algorithm is being proposed as a real-time guidance scheme the computational time that was needed to generate the entire trajectory by each method is presented in Table 7.2. While none of the codes have been optimized for computational efficiency, the use of quadratures does decrease the time needed to solve the launch problem in comparison to the shooting method. It should be noted that the flight time is approximately the same as the cpu time for the first-order approximation methods and that the shooting method was given a good initial guess (nearly converged) of the un- knowns. As expected, the zeroth-order analytic solution was found extremely 82 quickly. The introduction of the aerodynamic caused a modest increase in the computation pulse functions into the method time.