You are on page 1of 6

International Conference on Control, Automation and Systems 2010

Oct. 27-30, 2010 in KINTEX, Gyeonggi-do, Korea

Guidance and Control System Design for Impact Angle Control


of Guided Bombs
Yongho Kim1 , Jongju Kim1 and Minsu Park1
1
Department of PGM Technology, Hanwha Corporation R&D Center, Daejeon, Korea
(Tel : +82-42-829-2792; E-mail: aero21, jjkim2261, parkmins@hanwha.co.kr)

Abstract: Impact angle control has been widely used in a variety of guided weapons. For anti-ship or anti-tank weapons,
the terminal impact angle is important for warhead effect. This paper deals with the guidance and control system to
impact a target with a desired impact angle for precision guided bombs such as JDAM. The guidance and control system
is composed of three-loop autopilot and impact angle control guidance loop. The impact angle control guidance is derived
by the solution to the linear quadratic optimal control problem. Nonlinear six degree of freedom simulations are carried
out to examine the performance of the system.

Keywords: Guided Bomb, Guidance Kit, Autopilot Design, Impact Angle Control Guidance

1. INTRODUCTION posed of three-loop autopilot and impact angle control


guidance. Once the guidace and control system was de-
Proportional navigation guidance (PNG) law has been signed, nonlinear 6-DOF simulations were carried out to
most widely used in various guided weapons since the evaluate the performance of the guided bomb.
1940s. PNG only requires a small amount of information
that can be easily obtained from seeker mounted on the 2. GUIDED BOMB DYNAMIC MODEL
weapons or the ground radar, such as the closing velocity
and the line-of-sight (LOS) rate of guided weapons [1]. 2.1 Physical Properties
Impact angle control guidance law provides the solu- A guided bomb consists of a tail section with inte-
tion to achieve a desired impact angle as well as zero ter- grated aerodynamic control surfaces and a guidance con-
minal miss distance. It has been used to increase war- trol unit, a stabilizing strake kit, and a warhead as illus-
head effect of anti-ship and anti-tank guided missiles in trated in Fig. 1. In this paper, the guided bomb utilizes a
particular. For anti-ship guided missiles, it is employed Mk-82 500-pound warhead. One of the most common
to provide more favorable attack situation avoiding self- air-dropped bombs in the world, the Mk-82 is an un-
defense measures, such as surface-to-air missile systems, guided, low-drag general-purpose bomb (dumb bomb),
ECM(Electronic Counter Measures) systems, and CIWS part of the U.S. Mark 80 series. The guided bomb phys-
(close-in weapon system) [2-3]. Also, to ensure a high ical properties based on Mk-82 are given in Table 1 [6].
kill probability (or a penetration capability of warheads),
the achievement of a proper impact angle is desirable for
anti-tank missiles.
In addition, impact angle control has been used to sat-
isfy the flight path angle constraints of the waypoints of
unmanned aerial vehicles (UAVs), or to reduce the miss
distance produced by navigation errors for ballistic mis-
siles, or to design a terminal guidance system for reentry
vehicles with a constraint on the body attitude angle at
impact [4].
In this paper, impact angle control is applied to an
air-dropped guided bomb which strikes stationary targets
such as underground facilities or bunkers. The most rep-
resentative example of guided bombs is the JDAM (Joint
Direct Attack Munitions). The JDAM guidance kit is
a low-cost guidance kit that converts existing unguided Fig. 1 Guided bomb
free-fall bombs into accurately guided ”smart” weapons.
JDAM equipped bombs are guided to their target by an
integrated GPS/INS Navigation System [5]. 2.2 Equations of Motion
Missile DATCOM, the aero prediction software, was In order to evaluate performance, as well as to develop
used to obtain aerodynamic coefficients needed for a 6- a guidance and control system, it has to create a suitable
DOF dynamic model. The dynamic model of the guided model that meets the specific properties of the guided
bomb is based on the configuration of the Mk-82 with bomb. The numerical simulation model employed in this
guidance kit. The guidance and control system was com- paper consists of six degree of freedom model typically

978-89-93215-02-1 98560/10/$15 ©ICROS 2138


[ ]
Table 1 Guided bomb physical properties 𝑝𝑑
𝐿 = 𝑄𝑆𝑑 𝐶𝑙 (𝑛, 𝛼𝑡 , 𝛿, Γ) + 𝐶𝑙𝑝 (𝑛, 𝛼𝑡 , 𝛿, Γ)
Parameter Value 2𝑉
[ ]
Weight, kg 241 𝑞𝑑
𝑀 = 𝑄𝑆𝑑 𝐶𝑚 (𝑛, 𝛼𝑡 , 𝛿, Γ) + 𝐶𝑚𝑞 (𝑛, 𝛼𝑡 , 𝛿, Γ) (6)
Total Length, m 2.22 2𝑉
[ ]
Diameter, m 0.273 𝑟𝑑
𝑁 = 𝑄𝑆𝑑 𝐶𝑛 (𝑛, 𝛼𝑡 , 𝛿, Γ) + 𝐶𝑛𝑟 (𝑛, 𝛼𝑡 , 𝛿, Γ)
Center of Gravity from nose, m 0.95 2𝑉
Tail Fin Span, m 0.383 where 𝐶𝑖 , 𝑖 = 𝑥, 𝑦, 𝑧, 𝑙, 𝑚, 𝑛 are the non-dimensional
stability derivatives, 𝐶𝑙𝑝 , 𝐶𝑚𝑞 , and 𝐶𝑛𝑟 are the damp-
Axial Inertia, kg-m2 2.0
ing coefficients, 𝑄 is the dynamic pressure, 𝑔 is the ac-
Transverse Inertia, kg-m2 49.8 celeration due to gravity, 𝑉 is the total air speed, and 𝑆
and 𝑑 are the reference area and length, respectively. The
utilized in flight dynamic modeling. A flat Earth approx- body aerodynamic force and moment coefficients in Eqs.
imation is used to define the inertial coordinate system. (5) and (6) are functions of local Mach number 𝑛, total
The body frame is defined by the conventional manner. angle of attack 𝛼𝑡 , control fin deflection 𝛿 and bank an-
The dynamic equations are written with respect to body gle Γ. The definition of these aerodynamic coefficients
coordinates with appropriate kinematic equations relat- can be visualized in Fig. 2. The non-dimensional coef-
ing body translational and rotational rates and with iner- ficients are usually obtained through linear interpolations
tial translational and Euler angle rotational rates, respec- using data from look up tables. The aero prediction soft-
tively. The equations of motion are provided in Eqs. (1) ware Missile DATCOM was used to generate the non-
through (4) [7]. dimensional aerodynamic coefficients based on the bomb
⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎡ ⎤ physical properties of Table 1 [8].
𝑢˙ 𝑋/𝑚 0 −𝑟 𝑞 𝑢
⎣ 𝑣˙ ⎦=⎣ 𝑌 /𝑚 ⎦−⎣ 𝑟 0 −𝑝 ⎦⎣ 𝑣 ⎦ (1)
𝑤˙ 𝑍/𝑚 −𝑞 𝑝 0 𝑤
⎡ ⎤ ⎛⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎞
𝑝˙ 𝐿˙ 0 −𝑟 𝑞 𝑝
⎣ 𝑞˙ ⎦=[𝐼]−1 ⎝⎣ 𝑀˙ ⎦−⎣ 𝑟 0 −𝑝 ⎦[𝐼]⎣ 𝑞 ⎦⎠ (2)
𝑟˙ 𝑁˙ −𝑞 𝑝 0 𝑟
⎡ ˙⎤ ⎡ ⎤⎡ ⎤
𝜙 1 sin 𝜙 tan 𝜃 cos 𝜙 tan 𝜃 𝑝
⎣ 𝜃˙ ⎦=⎣ 0 cos 𝜙 − sin 𝜙 ⎦⎣ 𝑞 ⎦ (3)
𝜓˙ 0 sin 𝜙 sec 𝜃 cos 𝜙 sec 𝜃 𝑟
⎡ ⎤ ⎡ ⎤⎡ ⎤
𝑥˙ c𝜃c𝜓 s𝜙s𝜃c𝜓 − c𝜙s𝜓 c𝜙s𝜃c𝜓 + s𝜙s𝜓 𝑢
⎣ 𝑦˙ ⎦=⎣ c𝜃s𝜓 s𝜙s𝜃s𝜓 − c𝜙c𝜓 c𝜙s𝜃s𝜓 − s𝜙c𝜓 ⎦⎣𝑣 ⎦(4) Fig. 2 Definition of the aerodynamic coefficients
𝑧˙ −s𝜃 s𝜙c𝜃 s𝜙s𝜃 𝑤
Note the following variable definitions utilized in the 3. AUTOPILOT AND GUIDANCE LAW
equation above: 𝑢, 𝑣, and 𝑤 are the translational velocity
3.1 Autopilot Design
components resolved in the body frame; 𝑝, 𝑞, and 𝑟 are
the components of angular velocity vector expressed in In this section, we present a design technique for the
the body frame; 𝜙, 𝜃, and 𝜓 are the Euler angles; 𝑥, 𝑦, pitch, yaw, and roll autopilot design. The pitch/yaw au-
and 𝑧 are the components of the inertial position vector topilot chosen for this study is the so-called three-loop
expressed in the inertial reference frame; 𝑚 is the guided autopilot shown in Fig. 3. This autopilot design was cho-
bomb mass; [𝐼] is the inertia matrix about the body axis. sen because it is actively being used in several existing
In Eq. (4), we use the shorthand notations, 𝑠𝜙 = sin 𝜙, weapon systems and it can easily determine the appropri-
𝑐𝜙 = cos 𝜙, and so on. ate gains based on a few specified autopilot performance
The applied forces 𝑋, 𝑌 , and 𝑍 and moments 𝐿, 𝑀 , parameters. As Fig. 3 shows, the three-loop autopi-
and 𝑁 are given by Eqs. (5) and (6). The total forces lot takes acceleration commands from guidance and out-
acting on the guided bomb in Eq. (1) comprise the aero- puts the commanded deflection angles that should be re-
dynamic force and the gravitational force. The applied quired in order to achieve the commanded accelerations.
moments appearing in Eq. (2) contain contributions from These tail control surface deflections produce a response
the steady aerodynamic moment and the unsteady aero- through the airframe as a pitch/yaw rate and as a longitu-
dynamic. The unsteady body aerodynamic moment pro- dinal/lateral acceleration.
vides a damping source for bomb angular motion. The transfer function relating the achieved accelera-
tion and the pitch rate to the fin deflection can be written
𝑋 = −𝑄𝑆𝐶𝐴 (𝑛, 𝛼𝑡 , 𝛿, Γ) − 𝑚𝑔 sin 𝜃 as
( ) ( )
𝑌 = 𝑄𝑆𝐶𝑦 (𝑛, 𝛼𝑡 , 𝛿, Γ) + 𝑚𝑔 sin 𝜙 cos 𝜃 (5) 𝐴𝑧 𝑠2 2𝜁𝐴𝐹 𝑠2
= 𝐾1 1 − 2 / 1 + 𝑠+ 2 (7)
𝑍 = 𝑄𝑆𝐶𝑧 (𝑛, 𝛼𝑡 , 𝛿, Γ) + 𝑚𝑔 sin 𝜙 cos 𝜃 𝛿 𝜔𝑧 𝜔𝐴𝐹 𝜔𝐴𝐹

978-89-93215-02-1 98560/10/$15 ©ICROS 2139


Table 2 Linearized aerodynamic parameters (M=0.8)
Altitude 𝑀𝛼 𝑀𝛿 𝑍𝛼 𝑍𝛿
0km -101.3 -166.8 -0.161 -0.108
2km -79.5 -130.9 -0.129 -0.087
4km -61.7 -101.5 -0.103 -0.069
6km -47.2 -77.7 -0.081 -0.054

Table 3 Autopilot gains (M=0.8)


Altitude 𝐾𝐷𝐶 𝐾𝐴 𝜔𝐼 𝐾𝑅
0km 1.031 -0.119 15.637 0.447
Fig. 3 Pitch and yaw three-loop autopilot system 2km 1.025 -0.150 14.574 0.623
4km 1.020 -0.192 13.277 0.864
( ) 6km 1.016 -0.248 11.785 1.331
𝜃˙ 2𝜁𝐴𝐹 𝑠2
= 𝐾3 (1 + 𝑇𝛼 𝑠) / 1 + 𝑠+ 2 (8)
𝛿 𝜔𝐴𝐹 𝜔𝐴𝐹
If we neglect actuator dynamics, the closed-loop transfer
√ function 𝐺𝑐 (𝑠) of roll autopilot system can be written by
where 𝜔𝐴𝐹 = −𝑀𝛼 is the airframe natual frequency,
𝜁𝐴𝐹 = 𝑍𝛼 𝜔𝐴𝐹 /2𝑀𝛼 is the airframe damping, 𝜔𝑧2 = inspection of Fig. 4 as
(𝑀𝛼 𝑍𝛿 − 𝑀𝛿 𝑍𝛼 )/𝑍𝛿 is the airframe zero, 𝐾1 = 𝐾𝑃 𝐻𝐼 𝐾𝑃 𝐿𝛿𝑟
𝑉 (𝑀𝛼 𝑍𝛿 − 𝑀𝛿 𝑍𝛼 )/𝑀𝛼 is the aerodynamic accelera- 𝐺𝑐 (𝑠) =
𝑠2 + (𝐾𝑃 𝐿𝛿𝑟 − 𝐿𝑝 )𝑠 + 𝐾𝑃 𝐻𝐼 𝐾𝑃 𝐿𝛿𝑟
tion gain, 𝐾3 = −𝐾1 /𝑉 is the erodynamic body rate (10)
gain, and 𝑇𝛼 = 𝑀𝛿 /(𝑀𝛼 𝑍𝛿 − 𝑀𝛿 𝑍𝛼 ) is the turning rate 𝜔𝑛2
time constant. The linearized aerodynamic parameters = 2
𝑠 + 2𝜁𝜔𝑛 𝑠 + 𝜔𝑛2
are usually calculated at a trim angle of attack and sum-
marized in Table 2 for the four different flight conditions. Since Eq. (10) can be considered to be a second-order
The autopilot gains 𝐾𝐷𝐶 , 𝐾𝐴 , 𝜔𝐼 , and 𝐾𝑅 must be system, we require that 2𝜁𝜔𝑛 = 𝐾𝑃 𝐿𝛿𝑟 − 𝐿𝑝 and
chosen to satisfy some designer-chosen criteria (or some 𝜔𝑛2 = 𝐾𝑃 𝐻𝐼 𝐾𝑃 𝐿𝛿𝑟 . Therefore, we can solve for the
specified performance parameters) and can be determined roll autopilot gains yielding
by an analytical design methodology. Here, the desired
2𝜁𝜔𝑛 + 𝐿𝑝 𝜔𝑛2
performance parameters are the time constant 𝜏 , damp- 𝐾𝑃 = , 𝐾𝑃 𝐻𝐼 = (11)
ing ratio 𝜁 , and open-loop crossover frequency 𝜔𝐶𝑅 of 𝐿𝛿𝑟 𝐾𝑃 𝐿𝛿𝑟
the flight control system. In this paper, the open-loop The roll control gains can be determined easily for given
crossover frequency of 50rad/s was selected as a design design parameters, the closed loop natural frequency 𝜔𝑛
goal because the crossover frequency should be no more and the closed loop damping ratio 𝜁. Also, the con-
than one third of the bandwidth of the actuator model as- troller gains, 𝐾𝑃 and 𝐾𝑃 𝐻𝐼 are scheduled according to
sumed to have a natural frequency of 150rad/s. The other the Mach number and altitude.
two design goals were to achieve the desired flight con-
trol system time constant of 0.3s and the damping of 0.7.
The equations that determine the four autopilot gains (at
a single flight condition) are omitted in this paper. These
equations can readily be obtained from multiple sources
[1,9].
In this paper, the three-loop autopilot gains were de-
termined at sea level, 2km, 4km, and 6km altitude and at
Mach number 0.6, 0.8, 0.95, 1.05, and 1.2. Table 3 shows
the autopilot gains obtained for each of the four altitudes
at Mach number of 0.8. We can see from this table that
the autopilot gain 𝐾𝐴 and 𝐾𝑅 increases by a factor of 2
to 3 in going from sea level to 6km altitude whereas 𝐾𝐷𝐶 Fig. 4 Roll autopilot system
and 𝜔𝐼 decreases.
The roll attitude of a guided bomb can be controlled
3.2 Impact Angle Control Guidance
by a simple bank angle autopilot as illustrated in Fig. 4.
The transfer function relating the roll rate to fin deflection Guidance is the very important component to accu-
can be shown to be Eq. (9). rately commanding a guided bomb to a target. The guid-
ance algorithm used in this paper is impact angle con-
𝑝 𝐿𝛿𝑟 trol guidance law. Fig. 4 shows the guidance geometry
= (9)
𝛿𝑟 𝑠 − 𝐿𝑝 to impact a stationary target and impact angle definition

978-89-93215-02-1 98560/10/$15 ©ICROS 2140


Now, let us consider following optimal control prob-
lem: Find 𝑎𝑐 , which minimizes 𝐽, defined by

1 𝑡𝑓 𝑇
𝐽= 𝑢 (𝜏 )𝑅(𝜏 )𝑢(𝜏 )𝑑𝜏
2 0
∫ (20)
1 𝑡𝑓 𝑎2𝑐
= 𝑑𝑡, 𝑁 ≥ 0
2 0 𝑡𝑁 𝑔𝑜

subject to Eq. (17) and the terminal constraints given by

𝐷𝑥(𝑡𝑓 ) = 𝐸 (21)

where
Fig. 5 Guidance geometry and impact angle definition [ ] [ ]
1 0 𝑧𝑓
𝐷= , 𝐸= (22)
0 1 𝑣𝑓
on vertical plane. The distance 𝑟 and the LOS(Line-Of-
Sight) angle 𝜎 from the missile to the target are defined Here, the time-to-go is defined by 𝑡𝑔𝑜 = 𝑡𝑓 − 𝑡 and 𝑅 is
as follows. the positive weighting function defined as 𝑅 = 1/𝑡𝑁
𝑔𝑜 .

𝑟 = (𝑥𝑡 − 𝑥𝑏 )2 + (𝑧𝑡 − 𝑧𝑏 )2 The state feedback solution for the optimal control
( ) problem defined above can be obtained as
𝑧 𝑏 − 𝑧𝑡 (12)
−1 ( )
𝜎 = − sin 𝑢∗ (𝑡) = 𝑅−1 𝐵 𝑇 𝐹 𝐺−1 𝐹 𝑇 𝑥(𝑡) − 𝐸 (23)
𝑟
where (𝑥𝑏 , 𝑧𝑏 ) and (𝑥𝑡 , 𝑧𝑡 ) are the position of the bomb where
and the target, respectively. We can find the rate of
𝐹˙ = −𝐴𝑇 𝐹, 𝐹 (𝑡𝑓 ) = 𝐷𝑇
change of the distance and the LOS rate by differentia- (24)
tion of the Eq. (12), obtaining 𝐺˙ = 𝐹 𝑇 𝐵𝑅−1 𝐵 𝑇 𝐹, 𝐺(𝑡𝑓 ) = 0
𝑟˙ = −𝑉𝑏 cos(𝛾𝑏 − 𝜎) Substituting Eqs. (19) and (22) into Eq.(24), we have
𝑉𝑏 (13) 𝑇
𝜎˙ = − [sin 𝛾𝑏 − cos(𝛾𝑏 − 𝜎) sin 𝜎] 𝐹 (𝑡) = 𝑒𝐴 (𝑡𝑓 −𝑡) 𝐷𝑇
𝑟 cos 𝜎 [ ]
1 0 (25)
where 𝑉𝑏 and 𝛾𝑏 are the velocity and the flight path angle =
𝑡𝑓 − 𝑡 1
of the bomb.
Since the maneuver acceleration of the guided bomb is
∫ 𝑡𝑓
perpendicular to the velocity vector, the angular velocity ˙ )𝑑𝜏
of the flight path angle can be expressed as 𝐺(𝑡) = − 𝐺(𝜏
𝑡
𝑎𝑐 ⎡ ⎤
𝛾˙ 𝑏 = (14) (𝑡𝑓 − 𝑡)𝑁 +3 (𝑡𝑓 − 𝑡)𝑁 +2
𝑉𝑏 ⎢ − − (26)
𝑁 +3 𝑁 +2 ⎥
=⎢
⎣ (𝑡 − 𝑡)𝑁 +2

Under the assumption that 𝑉𝑏 is constant and 𝛾𝑏 and 𝜎 are
𝑓 (𝑡𝑓 − 𝑡)𝑁 +1 ⎦
small, after some algebra, we can linearize Eq. (13) as − −
𝑁 +2 𝑁 +1
𝑧˙𝑏 = 𝑉𝑏 𝛾𝑏 (15)
Substituting Eqs. (25) and (26) into Eq. (23) and ap-
The final conditions for the vertical position and the flight plying some algebra, we finally have the impact angle
path angle are specified as follows control guidance law for a lag-free systems [10-11] given
by
𝑧𝑏 (𝑡𝑓 ) = 𝑧𝑓 = 0, 𝛾𝑏 (𝑡𝑓 ) = 𝛾𝑓 = 𝜎𝑓 (16)
where 𝑡𝑓 denotes the time of flight, and 𝑧𝑓 and 𝛾𝑓 are the 𝑉𝑏
𝑢∗ (𝑡) = − [−𝑁𝜎 𝜎 + 𝑁𝛾 𝛾𝑚 + 𝑁𝑓 𝛾𝑓 ] (27)
desired terminal constraints. 𝑡𝑔𝑜
Eqs. (14) and (15) can be simplified further by letting where
𝑣(𝑡) = 𝑉𝑏 𝛾𝑏 (𝑡). Then, the state-space representation of
Eqs. (14) and (15) is given by 𝑁𝜎 = (𝑁 + 2)(𝑁 + 3),
𝑥˙ = 𝐴𝑥 + 𝑏𝑢, 𝑥(0) = 𝑥0 (17) 𝑁𝛾 = 2(𝑁 + 2),
𝑁𝑓 = (𝑁 + 1)(𝑁 + 2).
where
[ ]𝑇 [ ]𝑇 and the most widely used and the simplest time-to-go cal-
𝑥= 𝑧 𝑣 , 𝑥 0 = 𝑧0 𝑣 0 , 𝑢 = 𝑎𝑐 (18)
culation is the range over the bomb velocity, i.e., 𝑡𝑔𝑜 =
and 𝑟/𝑉𝑏 . Note that the parameter 𝑁 is the guidance gain
[ ] [ ]
0 1 0 chosen by the designer. For 𝑁 = 0, Eq. (27) becomes
𝐴= , 𝐵= (19)
0 0 1 pure energy optimal guidance law in [12].

978-89-93215-02-1 98560/10/$15 ©ICROS 2141


4. SIMULATION RESULTS Table 4 Initial conditions for 6-DOF Simulations
Parameter Value
In this section, we show the results of the nonlinear
6-DOF simulations. These simulations are carried out Bomb Position (0, 0, -6,000m)
using the Matlab Simulink model based on the system Target Position (10,000m, 1,000m, 0)
description given in the previous sections. The Simulink Bomb Velocity 250m/s (M=0.8)
model includes the WGS-84 gravity model and the ISA
Launch angle 0deg
atmosphere model. The autopilot loop is activated one
second after release to ensure safe separation from the Impact angle -30, -40, -50, -60, -70deg
aircraft, and then guidance loop is activated three seconds
after. No error sources considered and ideal measurement
6000
states are used for guidance.
We first perform the simulations for the comparison

Altitude(m)
4000
according to the guidace methods before examing the
simulation results for various impact angles. Fig. 6 and Unguided
2000
represent the history of the flight trajectory and the flight PNG
path angle regarding three different methods; unguided, IAC
0
PNG, and IACG. The maximum range and impact angle 0 2000 4000 6000 8000
without guidance for the initial conditions shown in Table Down Range(m)

4 result in a range of 8.49 kilometers and an impact angle 0

Flight Path Angle(deg)


of -57 degrees, respectively. Under the same range, PNG
−20
has an impact flight path angle of -50 degrees whereas
IACG matches the spcification of -80 degrees at the end −40
of the flight. Fig. 7 shows the differences in the guid- Unguided
−60 PNG
ance command and the angle of attack between PNG and IAC
IACG. −80
0 5 10 15 20 25 30 35 40
Figs. 8 shows the flight trajectories for various im- Time(sec)
pact angles. The guidance law achieve near zero miss
distance since no error source are considered. Fig. 9 Fig. 6 Comparison of flight trajectory and flight path
shows the flight path angles for various impact angles. angle according to guidance method
The desired terminal impact angles from -30 degrees to
-70 degrees are all met quite effectively. Although not
Guidance Command, A (m/s2)

30
given here, IACG can’t guarantee performance for impact Unguided
z

angle of more than 80 degrees. Because of an airframe 20 PNG


with a small g-limit capability, the limits of the maximum IAC
10
launch range and the terminal impact angle are inevitable.
For very low angles less than 30 degrees or very high 0
angles approaching 90 degrees, only a small increase in
−10
weapon effectiveness is gained with a large reduction in 0 5 10 15 20 25 30 35 40
Time(sec)
launch range.
Fig. 10 presents the time histories of the guidance 20
Angle of Attack, α(deg)

command for various desired impact angles. As can ob-


served form Fig. 10, the acceleration requirements are 0
dependent on the desired impact angles and the guid-
ance command tends to blow up as the guided bomb ap- Unguided
−20
PNG
proaches the target. These tendencies also make the total IAC
angle of attack sharply inceasing as shown in Fig. 11. −40
0 5 10 15 20 25 30 35 40
During the impact phase which is the last one second of Time(sec)
flight, the pitch attitude is actively controlled, to zero the
total angle of attack. This is done to align the warhead Fig. 7 Comparison of guidance command and AOA
longitudinal axis to the velocity vector to prevent war- according to guidance method
head breakup and to maximize the warhead effectiveness.
LAB Simulink model. The results of simulations indicate
that the proposed system can guide effectively the bomb
5. CONCLUSION
onto a designated target for the given initial conditions
In this paper, guidance and control system has been and various desired impact angle contitions.
designed for controlling terminal impact angle of the air- In general, for a given initial condition, the range of
dropped bomb. For performance analysis of the system, impact angle contollability is limited by the acceleration
the 6-DOF simulations are carried out using the MAT- capability of the airframe, and the time-to-go. Because

978-89-93215-02-1 98560/10/$15 ©ICROS 2142


this study hasn’t considered error sources, further study
40
will be necessary to perform the simulations containing
error sources such as Monte Carlo simulations. In addi- 30
tion to more accurate time-to-go calculation method, the
20
advanced guidance law to resolve constraints of guidance

Angle of Attack, α(deg)


performance due to the limited maneuverability and to 10
increase a range will be needed.
0 σ =−30°
f
σf=−40°
6000 −10
σf=−30° σ =−50°
f
σf=−40° σ =−60°
5000 −20 f
σf=−50° σf=−70°
σ =−60° −30
4000 f 0 5 10 15 20 25 30 35 40 45
σf=−70° Time(sec)
Altitude(m)

3000 Fig. 11 Angle of Attack for various impact angles

2000
REFERENCES
1000
[1] P. Zarchan, Tactical and Strategic Missile Guid-
ance, 5th ed., AIAA Inc., 2007.
0
0 2000 4000 6000 8000 10000 [2] R. K. Jung and Y. D. Kim, “Guidance Laws for
Down Range(m)
Anti-Ship Missiles Using Impact Angle and Impact
Fig. 8 Flight trajectories for various impact angles Time,” AIAA Guidance, Navigation, and Control
Conference, August 2006.
[3] J. I. Lee, I. S. Jeon, M. J. Tahk, “Guidance Law
to Control Impact Time and Angle,” IEEE Transac-
0
tions on AES, Vol. 43, Issue 1, 2007.
−10 [4] M. Kim, K. V. Grider, “Terminal Guidance for Im-
pact Attitude Angle Constrained Flight Trajecto-
−20 ries,” IEEE Transactions on Aerospace and Elec-
Flight Path Angle(deg)

−30
tronics Systems, Vol. AES-9, No.6, 1973.
[5] Kevin Wise, “Adaptive Flight Control of a Sen-
−40 σf=−30° sor Guided MK-82 JDAM,” Aerospace Control and
σf=−40° Guidance Systems Committee, 1998.
−50
σf=−50° [6] L. V. Krishnamoorthy, D. R. Kirk, R. Glass, “An
σf=−60° Aerodynamic Database for the Mk 82 General
−60
σf=−70° Purpose Low Drag Bomb,” DSTO-TR-0554, DoD
−70
0 5 10 15 20 25 30 35 40 45 Weapon Systems Division, 1997.
Time(sec) [7] R. C. Nelson, Flight Stability and Automatic Con-
Fig. 9 Flight path angles for various impact angles trol, Mc-Graw Hill, 1998.
[8] W. B. Blake, “Missile DATCOM User’s Manual -
1997 Fortran 90 Revision,” USAF, 1998.
[9] F. W. Nesline and M. L. Nesline, “How Autopilot
30 Requirements Constrain the Aerodynamic Design
σf=−30°
of Homing Missiles,” Conference Volume of 1984
20 σf=−40°
σf=−50°
American Control Conference, SanDiego, CA, June
Guidance Command, Az(m/s2)

σf=−60°
6-8, 1984.
10
σf=−70°
[10] C. K. Ryoo, H. Cho, and M. J. Tahk, “Time-to-
0
Go Weighted Optimal Guidance with Impact Angle
Constraints,” IEEE Transaction on Control System
−10
Technique, Vol.14, No.3, May 2006.
[11] C. K. Ryoo, “Impact-Angle-Control Guidance
−20 Laws for Maneuvering Targets,” KSAS Spring Con-
ference, November 2006.
−30 [12] C. K. Ryoo, H. Cho, and M. J. Tahk, “Optimal
0 5 10 15 20 25 30 35 40 45
Time(sec) Guidance Laws with Terminal Impact Angle Con-
straint,” Journal of Guidance, Control, and Dynam-
Fig. 10 Guidance commands for various impact angles
ics, Vol. 28, No. 4, 2005.

978-89-93215-02-1 98560/10/$15 ©ICROS 2143

You might also like