You are on page 1of 13

A

NEW LOOK

AT CLASSICAL VERSUS MODERN HOMING MISSILE GUIDANCE

F. Willim ,Nesline and P a u l Zarchan Raytheon Company, Missile Systems Division Bedford, Mastachusetts 0 1730
Abstract Modern guidance systems are generally accepted to yield better performance than classical proporticmal navigation systems. However, it is not always recognized that this better performance carries with it certain costs in inproved components or additional instruments. This paper compares a modern guiasnce system, MGS, to a classical proportional naviI V , homing misslle guidance system in gationai, P terms of perforuance, robustness, and ease of implementation. Q aan titative first order m i s s distances are compared to show that MGS has the smallest miss if component tolerances can be met, but a s component tolerances or measurement e r r o r s degrade, M G S degrades faster than P N until, at relatively large component or measurement erl-ors, P N has less miss distance than MGS. Introduction During the 1960's modern control theory was used in theoretical studies of closed form guidance laws for interceptor missiles. I t was shown that P N was an optimal soilltion to the linear guidance problem in the sense of producing zero m i s s distance for the least integral square control effort with a zero lag guidance system in the absence of target maneuver. ( 1 ) Aithough the reasons for the popularity of PN had nothing to do with that analysis, tbis important result gave credibility to the use of modern zontrol theory as a tool that many analysts have used to derive missile guidance laws. ( 2 P 3.4) Although much has been written concerning the mathematics of guidance, little, if any, has appeared in the open literature concerning the practical implementation of a modern guidance system, MGS of performance and implementation. Both guidance philosophies are reviewed and typical implementations are discussed. Finally both methods of guidance are comp~red in terms of performance and sensitivity to errors in implementation. Proportional Navigation

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

If two bodies are closing on each other they will eventually intercept i f the line of sight between the two does not rotate. Proportional navigation, PN, i s a method of guidance in which the m i s s i l e acceleration is made proportional to the line of sight rate. The geometry of a r , idealized intercept in which the missile and target are closing on each other at constant speed i s shown in Fig. 1. Here movement of the missile and target causc t h e line of sight to rotate through a small angle, A , indicating a differential displacement, y , between target and missile perpendicular to the reference. The PN guidance law is an attempt to mechanize an acceleration command, nc, perpendicular to the line of sight according to

where A ' is the effectjve navigation ratio, Vc is the closing velocity, and A i s the line of sight rate.

Proportional navigation has been in use for over three decades on radar, T V , and I R homin missile systems because of its effectiveness. ( 5 , 6 Although P N was apparently known b y the German scientists at Peenemiinde, no application using PN was reported. ( 7 ) I t was first studied by C. Yuan and others during World War I1 at the RCA Laboratories under the auspices of the U S Na-j; r ( g ) , it was extensively studied by Bennett 3nd Matthews at Hughes Aircraft and implemented in a pulse radar system(9), and it was fully developed b y H . Rosen and M. Fossier for a continuous wave radar system at Raytheon Company. The latter development included a closing velocity multiplier to compensate the guidance law dynamically in flight for changing engagement geometry. After World War 11, the U S work on P N was declassified and first appeared in the Journal of Applied Physics. ( 10)

..

Figure 1

Intercept Geometry

..

The purpose of this paper is to compare both classical and modern methods of guidance in terms

Copyright 1979 by F .W. Nesline and P. Zarchan with release to AIAA to publish in all forms.

The effective navigation ratio determines both the trajectory and acceleration history of the missile. For a zero lag guidance system, PN will result in zero miss distance [ y ( t ~ ) due to heading error =0] or target maneuver for any N' (assuming infinite missile accelex ation capability) This phenomenon i s clearly demonstrhted for the head-on case in the normalized trajectories shown in Fig. 2. Although relative tar get-missile iisplacement during the flight increases with decreasing ? I t , all flights result in zero miss distance. The effective navigation ratio

TARGET-WSGILES W A h n T l W W t i i

TARGET MANEUVtir

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

Figure 2

Typical Proportional Navigation Trajectories applications, o r measured by a doppler radar, a s in radar homing applications. The resulting acceferation command which is proportional to the line of sight rate estimate is applied to an acceleration utopilot that moves wing or fail control stlrfaces so as to develop the commanded acceleration.

m i s s distance Normaliz 3d missile acceleration histories due to both disturbances are shown in Fig. 3. Here missile a~celeration i s monctorlirally decreasing (except for N t = 2 ) for a heading e r r c r disturbance and monotonically increasing for a target maneuver disturbance. Figure 3 also shows that increasing N' minimizes the maximum acceleration due to target maneuver but maximizes the maximum acceleration due to heading error. In practice the navigation ratio is held fixed with acceptable values ( 3 < N ' C 5) determined by noise, radome and target maneuver considerations.

Other guidance concepts such as those used in modern guidance systems can best be w.derstood by studyink proportional navigation. I t can be observed from Fig. 1 that PN is mathematically equivalent ;0

A typical implementation of a P N guidance system is shown in Fig. 4 where an inertialiy stabilized seeker is used to measure t h e boresight e r r o r , E This signal, which is proportional to the line of sight rate, is low pass filtered to obtzin an estimate of the lirw of ight rate. The time constant, TN , of the noise filter r .In be fixed, as in this irnplementation, or time-varying to account for the range dependence of t\e measurement noise. The closing velocity c - either be estimated, as in 1R

.
The espression in the brackets of E q . ( 2 ) represents the miss distance that would result (in the absence of target maneuve-) i f the missile made no further ; the zero corrective accelerations and is referred to a effort miss, ZEM. Therefore P N can be thought of

MISSILE ACCELERATIONDUE TO HEADING ERROR

MISSILE ACCELERl

TIME

=F

Figure 3

Tvpical Proportional Navigration A c c e l ~ r a + i + -U G - f

--: me

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

Figure 4

P r o p o r t i o n a l Navigation Guidance S y s t e m Modern Guidance

PN

as a guidance law in which acceleration commands a r e issued inversely proportions; t o t h e square of time-to-go and directly proportional to ' h e ZEM. If t a r g e t m a n e u v e r , nT, is c o n s i d e r e d . t h e Z E M change3 and a new guidance law known a s augmented p r o p o r t i o n a l navigation, APN, r e s u l t s

N'
t

7 -[ y
go

2nTtgo2~

(3'

This guidance law is compared to P N , i n ,rms of trajectory and acceleration histories, for the case of a m a n e u v e r i n g t a r g e t with t h e r e s u l t s d b p l a y e d : laws achieve zero in Fig. 5. Although l ~ + f guidance miss distance, the tl L ' L :torr and acceleration r s e n t . The informatirn conhistories a r e vastly d cerning t a r g e t maneuver enabies A P N guidallce to u s e up l e s s a c c e l e r a t i o n capability than 9 N while keeping it c l o s e r t o an i n t e r c e p t c o u r s e . In addition t h e A P N a c c e l e r a t i o n h l s t o r y i s msnotonically d e c r e a s i n g unlike t h e monotanically i n c r e a s i n g h i s t o r y of P N . It can be shown that f a r N' = 3 A P N i s o p t i m a l in the s e n s e t h a t it achizves z e r o m i s s d i s t a n c e utlhaing t h e l e a s t i n t e g r a l <,quare c o n t r o l effort ( s t e Appendix C ).
TARGET-MISSILE SEFAhATION DUE TO TARGET MANFUJER

In a mc inrn guidance system, t h e ZEM is modified to take into account target maneuver ~ n d missile guidance system dynamics. If t h e guidance system dynamics can be r6presented by a first o r d e r transfer function, with bandwidth w , either modern control theory (11) o r the Schwartz Inequality ( s e e Appendix D ) cap b e used in deriving a guidance law wEich drives t h e miss distance to zerq vhile minimizihg the integral of the s q u a r e of t i a t accelera tion [ y ( t F ) = Osubject to minimizing law can be written a s

i , '

n c 2 d t ] . This

MISS11 E ACCELERATION DUE TO TARGET MANEUVER

TIME

tf

Figure 5

- Guidance Method C o m p a r i s o n

where

Appendices A and B for derivations vra Wiener and Kalrnan filter formulations). Thus the characteristic frequency of the filter increases with increasing process noise and decreases with increasing measurement noise.
A typical implementation of a modern guidance system appears in Fig,. 6. Here the line of sight angle is reconstructed from s seeker measurement of the boresight error and by integrating the rate gyro measurement of the seeker dish rate. This angle is then converted to relative target-missile position, y*, by the rnultiplicatiun of the r a g e measurement. T i n e signal is then sent through the Kalman filter i n order t o obtain estimates of the necessary states for the implementation of the modern guidance law. These states are multiplied by control gabs, which a r e funetiana,of the estimated ti&e to go and autopilot bmdwidth, in order t o genes ate an acceleration command, This c o m m a n d is applied to an a-eeleration autopilot in order t o develop the commanded acceleratim.

The expression within the brackets of Eq. (4) is the ZEM and Eq. (6) shows that the effective
navigation ratio is tirne-varying. The~aefoze this guidance law is a form of APN with an extra term to account for guidance system dynamics a1 i a timevarying navigation ratio. This guidance law, unlike P N , requires inbormat ion concerning time-lo- go, t , , , guidance system bandwidth, o , and achieved rni&ile acceleration, n~ Range measurements and additional filtering are required to estimate tgo and a c c ? k r o m e t e r measurements must now be fed into the guidance sys tern. The states required for the implementation of this guidance law ( y , +, n ~ can ) not be obtained from a simple low pass filter as was done s-ith P N , but must be e stimatsd,

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

A simple form of a Kalman estimator can be derived by considering the two most importart stochastic disturbances in a guidance s y s t e ~ , , random target maneuver and glint noise, T resultir e; Kalman filter is stationary and represented transfer function

In summary the implementation o f MGS requires some additional information which is not required by a dasaical PN guidance system. Estimates of range, measurement and process noise statis t i c s are needed for the %npl.emen%.ation of the Kalman filter whde estivates of time-to-go , guidance system bandwidth <.admissiie acceleration are needed for the implementation of the guidance l a w . Performance Comparison

Both classical and modern guidance zethods are


now compared in terms of performance as measured

with characteristic frequency, w

' given by

where r P , and b, are estimates of the spectral density l e v e l s o p t h e target rnaneu-. cr process noise and glint meec~urernentnoise respectively (See

by the rms (mot mean square) m i s s distance. The comparison is made utilizing the linearized, but realistic model of the kinematic homing loop shown in Fig. 7. Here autopilot dynamics are represented by a first order t r a n s f ~ .function and o d y 2-2 two most important s t o c h a s ~ c error sources are considered, namely glint noise and random target maneuver. The seeker, noise filter and guidance dynamics have been previously presented in Figs. 4 and 6. In this case the Kalman filter of MGS is optimal since it is ;wrfectly matched to the wrealworldM in that is has an exact dpnmical m o d d of the system along with pr f=Lectknowledge of the naeasurlerment and process

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

Figure 7

- Kinematic Homing Loop


then the influence of scale factor errors, C, or .MCS system performance can be investigated. Fig. shows t h a t system instabilities result if the scale factor falls below 0.6. A s before, PN performa i is not sensitive to this e r r o r source. In this for miss distances below 4.3 f t , only MGS can .I et the requirements if the scale factor is greater than 0.6. If the required miss distance is g r-eat e r than 4.3 f t , P N can always meet thc req&ement, but MGS can only meet the specification if the r ale factor is greater than 0.6. In summary, the impbementation of MGS places reqtiirements on allowable e r r o r s in dynamic modeling

noise statistics. With this methodology any deterioration in MGS performance will be caused solely by the guidance l a w . Imbedded in the hGS guidance law is a dynamical m o d i l of the actual system. If this model is inaccurate or if the tgo estimate is lacking or inaccurate, M G S performance degrades. If t h e estimated time to go, f 0, is considered to be a simple funct;on of tgoa s skown in E q . 9,

t h e n the influence of bias errors, B , and scale factor errors, A , on MGS system performance can be investigated Typical miss distance results, shown in Fig. 8, indicate that errors in result in rapid performance dekradation of MGS!*I~ fact, ~iegative bias e r r ~ r s lead to guidance system instabilities of MGS. P N performance, superimposec on F i g , 8 is not sensitive t o these e r r o r s . F o r this example P N achieves a m i s s of 4. 3 ft. Therefore if the required m i s s distance was l e s s than 4. 3 ft, only MGS could meet that specification and then only if bias e r r o r s could be kept below 0. 2 s and scale factor e r r o r s were between 0.68 and 1.35. If t.he required m i s s were greater than 4. 3 ft, P N could meet the specification, but MGS could only m e e t the specification i f bias and s c a l e factor e r r o r s could be kept below the values of t h e curves in Fig. 8. Of course, P N does not u s e t h e s e quantities at all and t h e r e f o r e i s not sensitive t o such e r r o r s . In summary, the implementation of MGS places requirements not only on the algorithm for calculating t but a l s o on the specialb-filtering n e e d e d to esk%ate range and range rate.

t'

Mcdern Guidance Sys tem, unlike Proportional Navigation, attempts to compensate for autopilot dynamics by the use of a dynamic lead t e r m in the missile acceleratioi~command. In order to implement this concept, MGS requires an accurate neasurement of the achieved missile acceleration and an estimate of the autopilot bandwidth. W A ~ .If this measurement is perfect and if the dynamic model within MGS is perfectly matched to the "real world", optimum performance can be obtained. However, if for example, we assume that missile acceleration is measured perfectly, but the estimate of autopilot
'
'

The nonhernispherical shape of the missile radome causes distortion of the incoming radar beam. A s the radar beam passes through t h e radome a refraction effect takes place and the net result is an e r r o r in the a ~ g l e of the apparent target. The radorne error slope, R , is a measure of the distortion taking place and is a function of the gimbal angle, among other things. (6) T h e guidance system designer attempts ;o specify the manufacturing tolerances and the limjts on tlre permissible variations of R, This error rrource i r particularly important a t high altitudes where t h e m i s s i l e turning rate t i m e constant, T , i s large. This t i m e constant in conjunction witg large radome refraction slopes can cause guidance system inatabiiity. It is, therefore, of considerable practical importance to s e e how P N and MGS performance degrade in the p r e s e n c e of radome slope e r r o r a . Typical high altitude p e r formance results f o r both guidance s y s t e m s a r e shown in Fig. 10. The r e s u l t s indicate t h a t t h e P N guidance system has m o r e of a tolerance t o radome e r r o r s than does MGS. In this regard P N is m o r e robust. Figure 10 indicates that the MGS implementation is more sensitive to negative radome slopes than positive slopes. In a practical deqjgn, the seeker stabilization loop gain could bp adjusted to bias the radome, thus i ~ s u r i n g only positive slopes. It is for this reason that the allowable radome slope range i s more of an important measure than the average radome slope. In fact, the allowable radome slope range is one important measure used in guidance svstem desipn to crnc!qy ~ - n q ~ f l - c + - - - - ; r , g tn1eran~-=nv the radome. The average rms miss distance due to a radome slope range can be calculated from the

'" . -

!A-

: --

rrvr-r

-- - -

qrrnrCJim +n

Frl

70

MGS

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

Figure 8

E r r o r s in Estimating Time To G o Seriously Degrade MGS Performance

t
V)

u ,
Y

af
u ,

I :

.1
Figure 9

E r r o r s in Estimating System Dynamics Can Lead t o MGS Instability

0 .OS .1 Normrlkod h d o m Slogl, Figure 10 Radorne E r r o r s hfluence System


-.06

Performance

information provided in Fig. 10. Typical r e s u l t s acceleration ratio for both guidance system showing t h e sensitivity of both guidance e ystcmrs . implementations, i t is evident that PN requires t o r a d o m e slope range i s displayed in Fig. 1 1. T h i s a larger acceleration advantage over the target than f i g u r e shows that when t h e r a d o m e is taken into con- MCS to achieve a specific m i s s distance. For sideration; MGS can only offer s u p e r i o r m i s s d i s example, to achieve an RMS m i s s of less than 7 f t , t a n c e p e r f o r m a n c e if t h e allowable radome slope PN requires a 5 to 1 acceleration advantage over the range is l e s s than 0.075. Otherwise P N yields target whereas MGS requires only a 2. 3 to 1 advans m a l l e r a v e r a g e r m s m i s s distances. Neverthe tage. This reduced accelerstion requirements less, if a r a d o m e slope range l e s s than 0.075 can extends the missile's zone of effectiveness against be met, MGS yields l e s s m i s s distance. maneuvering targets and is the major advantage of MCS over P N Finally, missile acceleration saturation, one of the most important guidance system nonlinearities Summary i n the performance comparison, is considered. Since MGS predicts intercept using estimates of A modern guidance system and a propnrtional target acceleration and measurements of missile navigation guidance system designed to meet the acceleration, it requires less acceleration than P N same miss distance specification yield different tn h i t mqn~vi~r~ri tn qr n- m t Tn Pin 1 3 . rrthC-n i-~!t;.:-clf~*i~:: : f ryb-.r;--+--=, -rid each subsvstem C r m s m i s s is plotted v e r s u s the m i s s i l e - t o - t a r g e t mUSt U l e e t a ClUielttilL be^ UA i r y w ~ ~ i ~ ~ ~ i i i a .

12

MGS

( 3)

Willems, G. , "Optimal Controllers for Homing tants, 'I U S Missiles with Two Time Cw.i Army Missile Commani, 3;;~ort No. RE-TR-69-20. Redstone Arsenal, Alabama, October 1969.

..

Lr. IV

cn
V)
H

h-

( 4)

Price, C F. , "Optimal Stochastic Guidance Laws for Tactical Missiles", The Analytical Sciences Corporation, Report No. TR- 170-2, Reading, Mass. , September 1971. Phillips, T. L , "Anti-Aircraft Missile Guidance" Electronic Progress published b y Raytheon Co. , March-April 1958, p p 1-5. Nesline , F W. , "Missile Guidance f o r Low Altitude Air Defenset', A U A Guidance and Control Conference, Palo Alto, CA, August 1978. Benecke, T. and Quick, A.W. (Eds) , "History of German Guided Missiles Development, " ACARD First Guided Missiles Seminar, Munich, Germany, A p r i l 1956. Yuan, C .L., "Homing and Navigational Courses of Automatic Target- Seeking Devices, 'I RCA Laboratories, Repcrt No. PTR- 12C, P r h c e t o n , N e w Jersey, December 1943. Bennett, R.R. and Mathews, W . E . , "Analytical Determination of Miss Distances for Linear Homin g Navigation Systems, Technical Memorandum No. 260, Hughes Aircraft Company, Culver City, California, March 1952. Yuan, C .L , "Homing and Navigational Courses of Automatic Target-Seeking Devices, " Journal of Applied Physics, December 1948.

( 5)

I I

(6)
.1
I

-05

,075

.15

NORMALIZED RADOME SLOPE RANGE

Figure 11
Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

- P N Allows G r e a t e r
Radome Design

Flexibility in

(7)

( 8)

(9)

( 10)
~ I S S l L LT O TARGET ACCELERATION R A T I O

Figure 12

- MGS Requires L e s s Acceleration Due


t o T a r g e t Maneuver
( 11)

modern guidance system imposes more severe requirements on radome refraction slope and on knowledge of t h e system dynamics, but i t does not require as much maximum missile normal acceleration to intercept an accelerating target. If t h e m i s s distance specification is extremely small, only MGS can do the job. Thus the additional instrumentation and subsystem requirements is the price that must be paid to meet severe miss distance requirements If the miss distance specification i s such that both M G S and P N can do the job, then instrumentation specifications can Ie relaxed in favor of more missile normal acceleration capability. When sufficient missile acceleration is available, PN offers t h e least stringent instrumentation requirements. Thus, if component tolerances can be met, MGS has the smallest miss distance, b u t as component tolerances or measurement e r r o r s degrade, the performance of MGS degrades f a s t e r than that of PN until at relatively large component o r measurement e r r o r s , P N has less miss distance than MGS.

Cottrell, R .G , "Optimal Intercept Guidance for Short-Range Tactical Missiles", AIAA Journal, Vol. 9, July 1971, pp 1414-1415. Fitzgerald; R J. , and Zarchan , P , "Shaping Filters for Randomly Initiated Target Maneuvers " , AIAA Guidance a n d Control Conference, Palo Alto, CA, August 1978.

. .

( 12)

(13) Wiener,

N., "Extrapolation, Lnterpolation and Smoothing of Stationary Time Serj esI1, MIT P r e s s , Cambridge, Mass., 1949.
Newton, G.C., Gould, L.A., and Kaiser, J . F . , "Analytical Design of Linear Feedback Controls", John Wiley and Sons, Inc., New York, 1957. Kalrnan, R . E . , and Bucy, R . , "New Results in Linear Filtering and Prediction, " Journal of Basic Engineering (ASME) , Vol 83 D , 1961, pp 95-108.

(14)

(15)

( 16)

Gelb, A . , e d . , "Applied Optimal Estimation, MIT Press, Cambridge, Mass. , 1974.

"

References
( 17)

(1) Bryson, A.E. and Ho Y . C . , "Applied Optimal Control, " Blaisdell Publishing Company, Waltham , Mass. 1969.
( 2)

Kliger , I. - " A Simple Derivation of Certain Optimal Control Laws I' , Raytheon Memorandum SAD- 1230, Bedford, ,Maas., Nov. , 1970.

Willerns, G , "Optimal Con trollers f o r Homing Missiles, " U . S. Army Missile Command, Report N o . RE-TR-68-15, Redstone Arsenal,
. " ! : t 3 2 , : ,

2 2 .'+-mhr.rf -

- C . n

-rn,, . .

236

Appendix A

- Wiener Gpthnal Filter


1herefore

The disturbances entering the guidance system are consilered to be white glint noise with spectral density @ly and random target maneuver. In this paper the maneuver is considered t r ~ be a step function whose initiation time is uniformly distributed over the flight time. I t can be shown( 12) that integrated white noise has the same autocorrelation function as this maneuver process. The optimal filter with transfer function, Ho, c z : ~ be derived by either Wirner or Kalman filter thcory. The G'riener filter formula?lon is based upon the diagram of Fig. A-1. The prablern is to find H, which will nkinimize the integral of the mean square error signal minimize /*eLd t ] ,
0

If we define

then Eq. (A-4) can be factored

The optimal transfer function, Ho, can b e found from the explicit solution of the Wiener-Hopf integral equation
Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

% :

where WS and WN are the spectral densities of the signal and noise, (WS+WN ) + represents that part which has all its poles and zeroes in the left half plane, (WS+WN)' represents that part which has all its poles and zeroes in the right half plane. The expression f 1+ is the component of ( 1 which has all its poles in the left half plane. In order to obtain [ ] + we expand ( ] in partial fractions and throw away all the terms corresponding to poles in the right half plane. F r o m Fig. A-1 the output spectral densities of t h e signal and noise, WS a n d can be expressed in t e r m s of the input specdensities. and QN, and the shaping network t r a n s f e r function a s

Therefore

Substitution of Eqs. (A-2 \, (A-,I 1 yields

( A - 7 ) and (A-8) into

Us = White Noise with Power Spectral Density Un = Whit. N o i u with Povnr Sp.ctral Wnsity

O,

q,,,

YT = Actual Targot Position VT* = Measured Target Position

VT
e

= =

Estimated Tarpat Position Error in Estimate

Appendix B

- Kalman Optimal Filter

The same transfer function can also be obtained by the Kalman formulation.
The state and measurement equations can bf: derived from the plant, shown in Fig. B- 1, and are

The expression in the brackets of Eq. ( A - 9 ) can be expanded by partial fractions yielding

( A - 10)

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

.nating those terms w i t h poles in the right half .ane leaves

which simplifies to

The Kalman filter equation i s

Substitution of E q . (.4- 12) into ( 4 - 9 ) yields t h e optimal transfer function

where the Kalman gains, 5 , are determined from the following matrix Riccati equations

I PROCESS I NOISE
I I

'I
I
I

I I I

PLANT
P

:
I

MEASUREMENT 0 NOISE I

us

-----+--I s ,
1
I

Y
4

1 - .

s
I

White Noise wi?h Power Spectral Density White Noise with Power Spectral Density Target Acceleration Target Rate Target Position Measured Target Position

= =

*,,,

tS

Y
Y, Y,*

= = =

Fig. B- 1 Kalrnan filter formulation.

Y OF THE OR1 GlNAL PAGE I S 1306

where

The transfer function between the position estimate output and the position measurement input can easily be obtained from Eq. ( B - i l l as
A

2s/wO+2 s 21 9 2

( E - 12)

7 -

YT -

y- T

1 + 2 S / 0 0 + 2 s 2 iw02

+ s3 I'J,

T i s is identical to the transfer function obtained by the Wiener filter approach.


, - is Recognizing that the covariance matrix, P symmetric, the scalar %quationsrepresenting the steady state solution (E = 2) can be written from {B-4) as

Appendix C - Matrix Approach to Modern Guidance Optimal guidance laws are generally derived on the basis of modern control theory. The linear model o f Figure C-1, in which missile dynamics are represented by a single lag, is used for the application of modern control theory to the guidance problem. I n this case we a r e interested in deriving a guidance law whish will achieve zero miss distance while minimizing the integral of t h e square of the guidance commands [minimize J rlLdt subject to v (t,)=01. 0 = The solution to minimizing the quadratic performance index
tf

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

After some algebra, the solutions to E q . (B-7) can L . e substituted into Eq. (B-5) yielding the steady state Kalrnan gains

subj ect to the linear differential equation

4=c?r+Cnc
\

(C-2)
( 1)

Defining

with zero miss distance f y ( t )=0 1 is well known F and is given b y

t h e gain matrix becomes

where R is obtained from the differential equation

The filter equations are obtained by substituting Eq. (B-10) into Eq, (B-3) yielding

4 .

.
h

1 0 .

-I'-

----

--

w-

--

---

REPROOU
&%-u--b

CE "15

I I

MISSILE DYNAMICS

4
1 I
!

LINEARIZED KINEMATICS

nc
Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

= = =

Commanded Missile Acceleration Achieved Missile Acceleration Target Acceleration

nL

Y
Y

= Relative Target

Y = Relative Missile Separation Fig. C- 1 Guidance Problem Formulation


The solutior, to E q . ( C - 4 ) is

Target -

Missile Rate

where

where T is the normalized time to go until intercept given by

If missile guidance dynamics are neglected (a+-), application of ~ ' ~ o ~ i t Rule a l d reduces the guidance law to

The integral appearing in E q becomes

. (C- 3)

therefore

This guidance law is augmented proportion a1 navigation with an effective navigation ratio of t h r e e .
Appendix D - Scalar Approach to Modern Guidance

Sdbstituting Eq. ( C - 8 ) into Eq. ( C - 3 ) yields the optimal closed loop guidance law

The Schwartz Inequality can also be used in the derivation of optimal guidance laws. [ For the zame problem considered in Appendix C , w e can express the system s t a t e vector of Eq. (C-2) at t h e terminal time according to the matrix superposition integral

where Q, is the fundamental matrix. Equation (D-1) actually represents a set of scalar equations for each of the different states. For the guidance problem, the first state is the object of our control and can be expressed as

In order to demonstrate the power of Eq. (D8) in solving for optimal guidance laws the example of Appendix C is reconsidered. The fundamental matrix can be found from Eq.. (C-5) to be

(D2)
For zero miss distance Eq. (D-'2) can be rewritten as and
Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

Application of the Schwartz Inequality yields

where t The integral of the square of the acceleration can be minimized by insuring that the equality sign of go =tF-t
(D- 12)

The integral of Eq. (D-5) becomes

E q . (D-5) holds. According to the Schwartz Inequality, this occurs when

Substitution of Eq. (D--6) into Eq. (D-3) yields

(D-13)

Substitution o f Eq. (D-13) into E q . (D-8)yields Therefore the acceleration is given by

where

and

A careful examination of the example used in both approaches shows the following equivalences

Thus modern control theory and the Schwartz Inequality can be used in the derivation of certain types of guidm-ce laws. The mathematical solutions of both approaches are in fact equivalent as can be seen by rewriting E q . (C-3) and (D-8).

Downloaded by CNPIEC (XI'AN BRANCH) on October 28, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.1979-1727

(D16)

You might also like