You are on page 1of 28

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/322309590

Development of a Full Envelope Flight Identified F-16 Simulation Model

Conference Paper · January 2018


DOI: 10.2514/6.2018-0525

CITATIONS READS
8 1,739

4 authors, including:

Marit Elisabeth Knapp Tom Berger


Santa Clara University U.S. Army Combat Capabilities Development Command Aviation & Missile Center
5 PUBLICATIONS   15 CITATIONS    45 PUBLICATIONS   331 CITATIONS   

SEE PROFILE SEE PROFILE

Mark B. Tischler
US Army Aviation Development Directorate - AFDD
157 PUBLICATIONS   2,686 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

High Speed Rotorcraft Modeling, Control, and Handling Qualities View project

UAS Guidance, Navigation, Control, and System Identification View project

All content following this page was uploaded by Marit Elisabeth Knapp on 01 March 2019.

The user has requested enhancement of the downloaded file.


Development of a Full Flight Envelope F-16 VISTA
Simulation Model from Closed-loop Flight Data

Marit E. Knapp
San Jose State University, Moffett Field, CA

Tom Berger †

Mark B. Tischler
U.S. Army Aviation Development Directorate, Moffett Field, CA
§
M. Christopher Cotting
Aaron Marcus ¶
USAF Test Pilot School, Edwards AFB, CA

A full flight envelope bare-airframe model of the Calspan NF-16D Variable-stability


In-flight Simulator Test Aircraft (VISTA) was identified from flight data. The unstable
nature of the aircraft required the control system to be engaged during flight and resulted
in highly correlated control surface measurements. To identify a bare-airframe state-space
model of the aircraft from flight data with the control system engaged (closed-loop), the
Joint Input/Output method was used to post-process frequency responses. A series of
discrete point models and trim data were used to develop the full flight envelope model
using a “stitched” model architecture. Employing a scaling method enabled the stitched
model to extrapolate to various loading configurations. This paper describes the identi-
fication methodology of the models with the control system engaged, the development of
the stitched model, and validation of the full flight envelope simulation model to flight data.

I. Introduction
he demands of highly maneuverable and responsive modern aircraft have led to configurations that are
T 1
aerodynamically unstable. Performing system identification to obtain accurate bare-airframe models
can be challenging when flight testing must be performed with a control system engaged, thus yielding
highly correlated control surface inputs. This paper explores system identification of the Calspan NF-16D
Variable-stability In-flight Simulator Test Aircraft (VISTA), see Figure 1. The VISTA operates with a
Variable Stability System (VSS) that provides a control architecture allowing the aircraft stability to be
varied with adjustable feedback gains that can produce controlled aircraft responses by suppressing aerody-
namic characteristics (e.g., unstable phugoid and spiral modes) that cause instability. Identification of the
aerodynamic contributions of the individual control surfaces and the overall aerodynamic characteristics of
the VISTA were of primary interest for building an accurate simulation model that behaves identically to
the aircraft. System identification of the VISTA was conducted in collaboration with the USAF Test Pilot
School. The project was to develop a full flight envelope “stitched model”2, 3 to accurately represent the
VISTA bare-airframe flight dynamics and trim characteristics across its flight and loading envelopes.
∗ Research Engineer
† Aerospace Engineer, Senior Member AIAA
‡ Flight Control Technology Group Leader, Senior Scientist, U.S. Army AMRDEC, Associate Fellow AIAA
§ Master Instructor of Flying Qualities, Associate Fellow AIAA
¶ Flight Sciences Technical Expert

Distribution Statement A: Approved for public release; distribution is unlimited.

1 of 27

American Institute of Aeronautics and Astronautics


Figure 1. Calspan NF-16D VISTA.

In many applications, it is desired to identify the open-loop characteristics of the aircraft or plant when the
system is operating with a controller engaged (closed-loop). The inherent problem with closed-loop data is
that it typically has less information about the open-loop system. When a system is operating closed-loop,
as the VISTA does, the measured responses may not excite the dynamics of the system adequately since
the feedback is meant to suppress undesired modes/characteristics to generate controlled responses (e.g.,
unstable systems). Additionally, output noise and process noise inside the feedback loop become correlated
with the bare-airframe input.3 The data correlation among dynamic variables of the control system and sup-
pression of important dynamics can result in inaccurate parameter identification with erroneous derivatives
that could be highly correlated due to the feedback action.3, 4

Specifically for fixed-wing applications, the aileron-to-rudder interconnect (ARI) present in the control sys-
tem poses a problem for bare-airframe identification because it directly correlates the aileron and rudder.3
When data is generated for system identification purposes of a dynamic multi-input system operating in
closed loop, the feedback causes correlations between the input variables,3, 4 making it difficult or impossible
to determine the input-output relationship associated with each of the individual inputs. This is especially
problematic for system identification from measured input-output data, as it is necessary that the measured
data contain accurate information about the modes of the system which requires determining the portion of
the response attributable to each input. As a result, it may not be possible to use a standard identification
approach5 alone to identify the open loop system parameters independently when a system is operating
closed-loop.

Even for plants that are asymptotically stable, opening the loop for identification may not be feasible due
to operational constraints. Identification of an open loop unstable plant, in which case opening the loop
for identification is prohibited, poses an even greater challenge. Alternate methods of system identification
for unstable systems operating closed-loop have been researched to accurately derive the open-loop system
dynamics.3, 6 In standard open-loop identification, the model is captured for the aerosurface command to
response behavior (y/δA ), as depicted in Figure 2 and demonstrated in Ref. 3. Generally, there are three ap-

Figure 2. Generic system identification block diagram.

2 of 27

American Institute of Aeronautics and Astronautics


proaches to closed-loop identification: Direct, Indirect and Joint Input/Output methods. The Direct method
treats the system as if it operates open-loop, where correlated inputs are either ignored or conditioned out.3
The Direct method cannot be used when correlation between relevant inputs is very large.

When correlation of inputs is high, alternate methods of identification can be used. Closed-loop identi-
fication using the Indirect approach incorporates the feedback (K) loop equations in the identification. To
extract the bare-airframe model from a closed-loop identification model with the Indirect method requires
accurate knowledge of the control system. Alternatively, the Joint Input/Output (JIO) method7, 8, 9 requires
no knowledge of the control system. Instead, the inputs and outputs of the bare-airframe are treated as
outputs of a system driven by uncorrelated piloted inputs. Herein, the JIO method was used to generate
open-loop frequency responses of the MIMO closed-loop system. This additional conditioning of the data
enables identification of bare-airframe derivatives from closed-loop flight data.

This paper covers the development of a full flight envelope stitched model for the VISTA. Section II re-
views the standard frequency-response system identification methodology, shown in Figure 3, along with
additional processing using the JIO method. Next, the point model identification process is presented in
Section III with example results of a point model for one of the flight conditions. A brief background on
the stitched model is given in Section IV followed by the integration of the identified point models to the
stitched model architecture. Each identified model and associated trim data was scaled to a common loading
configuration for the stitched model integration, presented in Section V. Lastly, the models were “stitched”
together to produce a continuous full flight envelope simulation model. Results and analysis of the full flight
envelope stitched model are presented in Section VI.

II. State-space Bare-airframe Identification Method


The system identification used in this research was the frequency response method, illustrated in the flowchart
in Figure 3 and implemented in the CIFER R
software tool.3 Frequency domain identification methods are
well suited to accurately characterize fixed-wing aircraft dynamics from flight data.3

Frequency Data Consistency Multi-variable


Aircraft & Spectral
Sweep Inputs
Reconstruction Analysis

Conditioned
Transfer-Function Frequency Responses
Modeling &
Partial Coherences

Frequency-Response +
Identification
Algorithm Identification
Criterion –

Initial Mathematical Model


Values Stability and Control Derivatives
and Time Delays

Sensitivity Analysis
&
Dissimilar flight Model Structure
data not used in Verification Determination
identification

Applications: FCS design, Handling-Qualities, Simulation validation

Figure 3. Frequency-response system identification method flowchart from Ref. 3.

The flight test plan of maneuvers and flight conditions for the aircraft is outlined in Section II.A. The primary
flight data used for identification were frequency sweeps and longitudinal static stability data. Before any
flight testing occurred, data consistency was checked, which is covered in Section II.B. System identification
requires an analyze of the input data for input cross-correlation. Correlation requirements and limitations
are discussed in Section II.C along with evaluation of the VISTA cross-correlation. Based on the level of

3 of 27

American Institute of Aeronautics and Astronautics


cross-correlation, the frequency responses from flight test data are conditioned. For this study, the frequency
responses were conditioned using the Joint Input/Output method, covered in Section II.D. The decorrelated
frequency responses from the JIO method enabled identification of linear state-space models of the VISTA
dynamics at each flight condition, which are integrated with trim data in a stitched model to produce a
continuous full flight envelope simulation model, as was previously done on the Calspan VSS Learjet and
described in Ref. 5.

A. Flight Test Overview


The flight tests provided a database of system identification data on the VISTA at various flight and loading
conditions, including piloted frequency sweeps, doublets, trim data, longitudinal static stability, and steady-
heading sideslip. Flight tests of the VISTA were conducted at Edwards Air Force Base by the USAF Test
Pilot School.

1. Aircraft Description
The aircraft used in this study is the Calspan NF-16D VISTA, a modified F-16 Fighting Falcon, single-engine
supersonic multirole fighter aircraft that carries an evaluation pilot and safety pilot, shown in Figure 1. The
lightweight fighter has relaxed static stability, with a cruise speed of 300 KCAS, maximum speed at sea
level of Mach 1.2 and a service ceiling of 50,000 ft. This unique aircraft has a custom digital flight control
computer. With the VSS engaged, the evaluation pilot flies the aircraft from the front cockpit, while the
safety pilot located in the back cockpit manipulates the VSS settings (e.g. freezing certain control surfaces).
Other modifications of the VISTA configuration include a heavy weight landing gear and a larger capacity
hydraulic pump and lines to accommodate the increased surface motions needed for in-flight simulation.

The VISTA is flown with the VSS, which operates symmetric and asymmetric horizontal tails and flap-
erons, rudder, and throttle control. The only surfaces not controlled by the VSS are the leading edge flaps,
which are scheduled with angle of attack. The VSS flight envelope is more constrained than the F-16 en-
velope, limited by the numerous safety trips that turn off the VSS and return nominal F-16 control law
augmentation in the event of exceedance of aircraft or aircrew safety limits. This study focused on the
operational flight envelope with the VSS engaged, which is limited to Mach 0.9 or 440 KCAS. The sensors
available for data collection on the VISTA measure pilots stick and throttle controls (δlat , δlon , δthrt ), ac-
tuator commands and corresponding surface deflections (δa , δe , δr , δf , δh ), air data (α, β, VT AS ), aircraft
states (φ, θ, ψ, p, q, r), and accelerometers (ax , ay , az ), all data signals were sampled at 64 Hz.

2. Flight Conditions/Test Points


The flight test plan was created based on the aircraft description/characteristics and guidance taken from
Tobias and Tischler2 in the selection of altitude and airspeeds of the identification flight conditions. Fre-
quency sweeps were performed in the pitch, roll, and yaw axes to produce frequency responses used to
identify the bare-airframe derivatives at each test point. A good example of a piloted longitudinal frequency
sweep is shown in Figure 4. In this research the frequency range and magnitude of the frequency sweeps was
especially important because the control system damped out the unstable aircraft dynamics. When possible,
piloted frequency sweeps were collected and long record lengths were desired to capture the low frequency
dynamics. The exception was identification of the symmetric flaperons, which could not be flown by the
pilot, requiring an automatic injected test input sweep instead.

4 of 27

American Institute of Aeronautics and Astronautics


θ [deg] q [deg/s] δe [deg] δlon [in]
2
0
−2
10
0
−10
20
0
−20
10
0
−10
10
az [ft/s2 ] ax [ft/s2 ] α [deg]

5
0
10
0
−10
0
−50
−100
0 20 40 60 80 100 120 140 160
Time [sec]

Figure 4. Example longitudinal axis frequency sweep time history.

Figure 5 shows the test flight conditions and maneuvers flown. At each test point used in model identification,
frequency sweep, doublet, longitudinal static stability data and steady heading sideslip data were collected.
Doublet maneuvers were performed in the pitch, roll, and yaw axes for model validation. Additional trim
data was collected during the Longitudinal Static Stability (LSS) maneuver, which produces 10 second trim
records at various airspeeds. The LSS maneuver flight data was used to augment the low-frequency iden-
tification of the longitudinal responses. Steady Heading Sideslip (SHSS) maneuver flight data was used to
verify the low-frequency identification of the lateral/directional responses.

Test point 3*, indicated with the pink diamond in Figure 5, was the anchor point for parameter identification
and was repeated in a heavy and light fuel weight configuration to enable calculations of the moments of
inertia to scale the stitched model.5, 3 Trim shot data were collected in 20 KIAS increments at 20 kft as
presented in Figure 5. Longitudinal and lateral/directional state-space models were identified at all of the
test points 1 − 5, indicated as the Model Identification points, covering the flight envelope shown in Figure 5.

5 of 27

American Institute of Aeronautics and Astronautics


80000
Flight Envelope
Ground Level at Edwards AFB
Sweeps, LSS, SHSS, Doublets for Model Identi-cation
Trim Shot Data
Full and Empty Fuel Tank Loading Con-gurations
60000
Doublets for Model Development

Altitude [ft]

40000

1 2 3$ 4 5
20000

0
0 0:2 0:4 0:6 0:8 1

Mach

Test points and maneuvers to cover full flight envelope (flight envelope depicted for the
Figure 5.
VISTA with VSS engaged).

B. Angular and Translational Data Consistency


As a precursor to system identification, it was essential to check for kinematic consistency and perform recon-
struction techniques to ensure that any sources of error were corrected for and that flight data measurements
were dynamically relatable.3 Angular and translational consistency were checked by reconstructing response
data with the assumption that small trim angles and perturbation motions would allow for use of linearized
equations. For example, attitude and attitude-rate frequency responses have identical dynamic information
needed for identification when the data is kinematically consistent,3 shown for pitch:

q = sθ (1)

The angular consistency check was performed by forming the measured attitude-to-angular rate frequency
response extracted from flight data. The consistency of the pitch rate and angle responses was evaluated
with a gain and time delay transfer function (TF) fit:
1q
Ke−τ s = (s) (2)

over the frequency range of 0.5 − 7.2 rad/sec, where the flight data is indicated in blue and the transfer func-
tion fit is in red in Figure 6. Any scale factor or time shift difference would indicate kinematic inconsistencies
between the responses. The resulting gain and time delay are 1.04 and 15.5 ms respectively, indicating good
kinematic consistency between the two signals. The results of the angular kinematic data consistency check
showed good agreement between the attitude and angular rate in both the time domain and frequency do-
main. Generally, the angular-rate response had higher coherence for the entire frequency range of interest.
The pitch attitude had higher coherence at low frequency so it was included for low frequency range in the
longitudinal identification.

For this application, checking the consistency of the feedback sensors was crucial so that time delays could
be incorporated into the identification. Figure 7 shows the angle of attack consistency check relation:

ẇ = sUo α (3)

6 of 27

American Institute of Aeronautics and Astronautics


where the transfer function fit of the frequency response over 0.4 − 8.1 rad/sec resulted in a gain and time
delay of 1.05 and 45.3 ms respectively. The gain indicated good agreement in the magnitude of the responses,
but the difference in time delay was due to additional filtering of the angle of attack response. The effective
time delay τα was found by evaluating the time delay as shown in Figure 7. The angle of sideslip β was also
filtered and had a time delay, resolved as stated above. Excellent fit between the rest of the measurements
ensured confidence in consistent flight data for the identification process.

20 20

Magnitude [dB]
Magnitude [dB]

10 10

0 0

-10 -10
q=3s -20 w=sU
_ o,
-20
100 TF -t Ke!=s = 1:04e!0:0155s (J = 2:82) 100 TF -t Ke!=s = 1:05e!0:0453s (J = 6:63)

Phase [deg]
Phase [deg]

50 50
0 0
-50 -50
-100 -100
1 1

Coherence [-]
Coherence [-]

0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

Figure 6. Pitch rate/attitude consistency check. Figure 7. Angle of attack/vertical velocity consis-
tency check.

The effective time delay on angle of attack response, τα cannot be neglected in the state-space model because
it is one of the outputs. This was taken into account by adding a sensor delay, represented with an equivalent
first-order Padé filter in the CIFER R
identification set-up. The delayed angle of attack αdelay that is fed
back to the controller is used in the identification with Equation 4.

αdelay −s + 2/τα
(s) = (4)
α s + 2/τα

The delayed angle of sideslip was used for identification with this same method.

C. Input Cross-correlation Evaluation


System identification from flight data requires that: control surfaces are not highly correlated.3 One of
the major problems with bare-airframe identification from closed-loop test data is that the controller can
introduce significant correlation between the control inputs. Ref. 3 studies how this control correlation
can degrade the quality of the identification. If the control loop system dynamics are only weakly excited,
then the feedback loop will only induce partial correlation and the Direct method for bare-airframe system
identification can be used to find accurate derivatives by using input at the aerosurfaces and multi-input
processing. However, when a high gain feed forward or feedback loop is present, as shown in Figure 8 for
the lateral/directional axis, and aerosurfaces may be highly or completely correlated, which would cause the
parameters to have large errors and uncertainties.3, 4 Although the ARI was turned off for flight testing, the
block diagram, shown in Figure 8, illustrates that when a lateral maneuver is performed, the rudder will
be excited by Kβ feedback. Hence, it was necessary to determine the extent of the aerosurface correlation
present on the VISTA.

7 of 27

American Institute of Aeronautics and Astronautics


Figure 8. Simplified block diagram of the VISTA F-16 lateral/directional axis.

The level of control input correlation was determined by coherence and autospectrum evaluation of the in-
puts. The coherence function is a measure of the portion of the output that is linearly attributable to the
input of a frequency response.3 The level of correlation between two signals (x,y) can be evaluated with the
2
coherence function, where low average coherence (i.e., (γx,y )ave ≤ 0.53 ) indicates low correlation. Figure 9
shows the cross-control coherence between asymmetric flaperons and rudder during a closed-loop sweep of
the rudder. The low average coherence over the entire frequency range in Figure 9 during the rudder sweep

1
./2a ;/r : Asym Flaperons /a
0:9
0:8
0:7
0:6
Coherence

0:5
0:4
0:3
0:2
0:1
0
10!1 100 101 102
Frequency [rad/sec]

Figure 9. Asymmetric flaperons coherence during frequency sweep of the rudder.

means the asymmetric flaperons were not heavily correlated with the rudder, indicating Direct MIMO iden-
tification method could be used.

In this study, the ARI on the VISTA flight controller was turned off, but the feedback correlates the rudder
and asymmetric flaperon surface motions. Figure 10 shows that during roll sweeps performed with the asym-
metric flaperons, the two control inputs are highly correlated, indicated by the high cross-control coherence.
Figure 11 shows that the rudder was heavily excited during the asymmetric flaperon sweeps, because its
power spectral density was within 10 − 20 dB of the swept aerosurface.3 During the asymmetric flaperon
roll sweeps, the rudder cannot be conditioned out with conventional multi-input processing because of the
high input correlation.3 As such, the control input correlation would compromise the bare-airframe identifi-
cation in the lateral/directional axis if the Direct method was used. The same analysis was repeated for the
longitudinal axis, where correlation was present between the symmetric flaperons and horizontal tail during
the symmetric flaperon sweep (horizontal tail moves due to α and α̇ feedback).

8 of 27

American Institute of Aeronautics and Astronautics


1 10
0:9 G/r ;/r
0 G/a ;/a
0:8
0:7 !10

Magnitude [dB]
0:6
Coherence

!20
0:5
0:4 !30

0:3
!40
0:2
!50
0:1 ./2r ;/a : Rudder /r
0 !60 !1
10!1 100 101 102 10 100 101 102
Frequency [rad/sec] Frequency [rad/sec]

Figure 10. Rudder coherence during asymmetric Figure 11. Comparison of rudder and flaperon
frequency sweep of the flaperons. aerosurface autospectrum during asymmetric fre-
quency sweep of the flaperons.

D. Joint Input/Output Method for Input Decorrelation


Based on high correlation between the aerosurface inputs, the Joint Input/Output method was used to gen-
erate frequency responses for system identification. The JIO method utilizes frequency responses from pilot
inputs-to-aircraft responses (y/δS ), and pilot inputs-to-correlated aerosurface inputs (δA /δS ) to reconstruct
aircraft frequency responses due to aerosurface inputs (y/δA ). Equation 5 expresses the matrix of frequency
responses
     −1
y(jω) y(jω) δ A (jω)
= × (5)
δ A (jω) δ S (jω) δ S (jω)
computed on a frequency-by-frequency basis. To evaluate the quality of the calculated frequency responses
and for system identification, it is important to access the coherence of the responses. From analysis of the
coherence of the individual frequency responses that are used to to produce frequency responses with the
JIO method, a weighted average coherence can be calculated and used for system identification. This study
uses the calculation for coherence described in Ref. 7.

1. Validation of Joint Input/Output Method


During the symmetric horizontal tail pitch sweeps, the symmetric flaperons were not correlated with the
horizontal tail, which made it the ideal test case because the flight data could be processed with the Di-
rect method and compared to data processed with the JIO method for the closed-loop test configuration.
Figures 12 and 13 show excellent agreement between the frequency response generated directly from (uncor-
related) aerosurface input-to-aircraft response and the frequency response generated using the JIO method.
Validation of the JIO method for the symmetric horizontal tail case gave confidence that the JIO method
was applied correctly and would produce accurate frequency responses for highly correlated inputs.

9 of 27

American Institute of Aeronautics and Astronautics


0 !20

Magnitude [dB]

Magnitude [dB]
!30
!20
!40
!40
!50

!60 !60
-90 -90
Phase [deg]

Phase [deg]
-180 -180
-270 -270
-360 -360
-450
1 -450
1
Coherence [-]

Coherence [-]
0:5 0:5
q=/e Direct Method r=/r Direct Method
q=/e Joint I/O Method r=/r Joint I/O Method
0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

Figure 12. Pitch rate response from Direct method Figure 13. Yaw rate response from Direct method
compared to the response generated with the Joint compared to the response generated with the Joint
Input/Output method. Input/Output method.

III. Point Model Identification Results from Flight Test Data


A. Identification Algorithm
Longitudinal and lateral/directional state-space math models were identified with the VISTA fight data.
The models are in body axes, and take the following form in CIFER R 3
:

ẋ = Ax + Bu(t − τ )
(6)
y = Cx + Du(t − τ )

The longitudinal model is given by Equation 7 and the lateral/directional model by Equation 8, where the
aerosurface time delays τ are not shown but were included in identification.

      
u̇ Xu Xw Xq − Wo −g cos(Θo ) u Xδe Xδf XδT  
 δe
ẇ   Zu Zw Zq + Uo −g sin(Θo )   w   Zδe Zδf ZδT  
     
= +   δf
 
  
 q̇  Mu Mw Mq 0   q  Mδe Mδf MδT 
δT
θ̇ 0 0 1 0 θ 0 0 0
     
q 0 0 1 0   0 0 0 (7)

α   0 1/Uo 0 0  u  0 0 0 
    

       δe
ax  Xu Xw X q 0 w X X XδT 
+  δe δf
  
 =    δf

az  Z Zw Zq 0  q  
 Zδe Zδf ZδT 
  
  u
 δT
  
u̇  Xu Xw Xq − Wo −g cos(Θo ) θ Xδe Xδf XδT 
    

ẇ Zu Zw Zq + Uo −g sin(Θo ) Zδe Zδf ZδT

δe symmetric horizontal tail stabilizer aerosurface (acts like elevator)


δf symmetric flaperon aerosurface
δT thrust input

10 of 27

American Institute of Aeronautics and Astronautics


      
v̇ Yv Yp + Wo Yr − Uo g cos(Θo ) v Yδa Yδh Yδr  
 δa
 ṗ   Lv Lp Lr 0   p   Lδa Lδh Lδr  
     
= +   δh

  
 ṙ  Nv Np Nr 0   r  Nδa Nδh Nδr 
δr
φ̇ 0 1 tan(Θo ) 0 φ 0 0 0
(8)
     
p 0 1 0 0   0 0 0 
 v

 r    0
 0 1 0   0
 0 0  δa
 p  
 
ay  =  Yv Yp Yr 0  + Y Yδh Yδr   δh
    
  r   δa
  
   
 β  1/Vtot 0 0 0   0 0 0  δr
φ
v̇ Yv Yp + Wo Yr − Uo g cos(Θo ) Yδa Yδh Yδr

δh asymmetric horizontal tail stabilizer aerosurface


δa asymmetric flaperon aerosurface (acts like ailerons)
δr rudder aerosurface

The derivatives of the body velocities (e.g., u̇) were reconstructed from inertial measurements for use in
the model identification from inertial data.3 The thrust input δT [lbs] was calculated from the measured
pilot throttle input δt [deg] and modeled with a second order system given in Equation 9.2

ωT2
δT = δt (9)
s2 + 2ζT ωT s + ωT2

1. Longitudinal Static Stability Data


For state-space identification, there is typically reduced frequency response coherence at frequencies low
enough to accurately capture the phugoid dynamics (evident by poor coherence at frequency under 0.3 rad/sec
in Figure 16). Therefore, directly identifying the u stability derivatives, associated with low frequency
phugoid mode, would result in large insensitivity values for those derivatives. Rather than identifying the u
stability derivatives in the frequency domain, identification of u stability derivatives can be improved using
the longitudinal static stability as shown herein.3, 2 This was also done for the identification of the fixed-
wing business jet in Ref. 5, and ensures that the speed derivatives will be consistent with the stitched model.5

Longitudinal static stability data were gathered using the stabilized point technique10 and used for the iden-
tification of the u stability derivatives. The aircraft was trimmed around the identification point (300 KIAS,
20 kft), throttle position was held constant, and airspeed was varied by climbing and descending. Airspeed
was stabilized at three points above and three points below the trim condition (approximately 10 kts apart).
Figure 14 shows the full maneuver time history. Highlighted in red on the time history data are sections
where the airspeed was stabilized. Trim pitch attitude θ, angle of attack α, horizontal tail surface deflection
δe , and throttle δt were averaged over each of the highlight sections. These trim points are plotted against
airspeed to determine the trim gradients with respect to airspeed as shown in Figure 15. The trim gradients
were used to determine stability derivatives (Xu , Zu , and Mu ) in the state-space model identification. This
is done by recognizing that in trim the states are constant, ẋ = 0. Then, the state-space equations can be
used to solve for the u derivatives as a function of the other derivatives and the trim gradients, as is shown
here for the Xu derivative:3

∆u̇ = Xu ∆u + Xw ∆w + (Xq − Wo )∆q − g cos Θo ∆θ + Xδe ∆δe + Xδf ∆δf + XδT ∆δT (10)

∆w ∆θ ∆δe ∆δf ∆δT


Xu = −Xw + g cos Θo − Xδe − Xδf − XδT
∆u ∆u ∆u ∆u ∆u
∆w ∆θ ∆δe ∆δf ∆δT
Zu = −Zw + g sin Θo − Zδe − Zδf − ZδT (11)
∆u ∆u ∆u ∆u ∆u
∆w ∆δe ∆δf ∆δT
Mu = −Mw − Mδe − Mδ f − MδT
∆u ∆u ∆u ∆u

11 of 27

American Institute of Aeronautics and Astronautics


For the longitudinal axes identifications the speed derivatives Xu and Zu were fixed, while Mu was set as
the constraint Equation 11.

800 6
U [ft/sec]

Trim data "3="u = -0.01111 deg / ft/s

3 [deg]
600
4
400
2
10
3 [deg]

60
0

w [ft/ses]
"w="u = -0.074059 ft/s / ft/s
!10 40

10 20
, [deg]

5 2:5

/e [deg]
0
2
4 "/e ="u = 0.0012908 deg / ft/s
/e [deg]

2 1:5

0 2

/f [deg]
"/f ="u = 0.00056088 deg / ft/s
4 1.8
/f [deg]

2
1.6
0
44
50

/t [deg]
"/t ="u = 0.001123 deg / ft/s
/t [deg]

43
40
30 42
0 50 100 150 200 250 300 350 400 500 550 600 650 700 750

Time [sec] U [ft/sec]

Figure 14. Longitudinal static stability time history. Figure 15. Longitudinal static stability selected
trim data.

B. State-space Model Identification Results (Full Fuel Tank at 300 KIAS)


The state-space model given in Equations 7 and 8 were identified using the frequency responses shown in
Figure 16 using the DERIVID tool in CIFER R 3
. Initial guesses for the model parameters were obtained
from transfer-function identification results and a priori estimates based on first principles.5, 11 The figures
show excellent agreement between the flight data and model, which is confirmed by the low individual and
average fit costs given in Table 1. The accuracy of the math model is quantified in the cost function, which
is calculated as a weighted sum of the magnitude and phase errors between the frequency responses and the
model responses (cost of J < 50 indicate excellent agreement3 ).

Table 1 shows that many of the outputs have duplicate measurements, like az and az2 , where two accelerator
measurements from different locations in the aircraft were used. This redundancy gave the software addi-
tional information to identify the state-space model. The identified parameter values as well as insensitivity
and Cramér-Rao bounds are given in Table 2. All identified parameters have good accuracy and are not
correlated with other parameters as indicated by their low insensitivities (I ≤ 10%) and Cramér-Rao bounds
(CR ≤ 20%). The zero value for Mw shows that the VISTA has no static stability (no static margin), as
expected for the relaxed static stability F-16. Eigenvalues presented in Table 3 show that at 300 KIAS with
a full fuel tank, the longitudinal axis is stable for the short term modes, it is slightly unstable for the long
term, phugoid mode. This process of identifying a model from frequency responses from the flight data was
repeated for the lateral/directional axis and at each flight condition. The eigenvalues from the 300 KIAS
lateral/directional axis flight identification are presented in Table 3. All of the lateral/directional modes are
stable for the 300 KIAS with a full fuel tank loading configuration.

12 of 27

American Institute of Aeronautics and Astronautics


q=/e ,=/e
!10 0

Magnitude [dB]

Magnitude [dB]
Frequency response Frequency response
!20 Model identi-cation !20 Model identi-cation

!30 !40

!40 !60
0 0
Phase [deg]

Phase [deg]
-180 -180

-360 -360

-540 -540
1 1
Coherence [-]

Coherence [-]
0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

a) b)

ax =/e az =/e
30 60
Magnitude [dB]

Magnitude [dB]
20
Frequency response Frequency response
Model identi-cation 40 Model identi-cation
10
20
0

!10 0
0 0
Phase [deg]

Phase [deg]

-180 -180

-360 -360

-540 -540
1 1
Coherence [-]

Coherence [-]

0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

c) d)

u=/
_ e w=/
_ e
30 50
Magnitude [dB]

Magnitude [dB]

20
Frequency response Frequency response
Model identi-cation 40 Model identi-cation
10
30
0

!10 20
0 0
Phase [deg]

Phase [deg]

-180 -180

-360 -360

-540 -540
1 1
Coherence [-]

Coherence [-]

0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

e) f)

Figure 16. a) q/δe b) α/δe c) ax /δe d) az /δe e) u̇/δe f ) ẇ/δe Longitudinal frequency response and
identification results for 300 KIAS, heavy configuration at 20 kft.

13 of 27

American Institute of Aeronautics and Astronautics


Table 1. Longitudinal Table 2. Longitudinal DERIVID results, 300 KIAS
DERIVID results, heavy loading configuration.
300 KIAS heavy
loading configuration.
Param. Value CR (%) Insens. (%)
Response Cost F-matrix
α/δe 35.51 Xu -0.002277a − −
Xw 0.04917 2.057 0.9349
q/δe 11.49
ax/δe 7.105 Xq 0b − −
ax2 /δe 16.6 −g cos(Θo ) -32.09a − −
a
az/δe 4.744 Zu -0.04708 − −
Zw -0.6184 2.632 1.024
az2 /δe 7.871
u̇/δe 47.95 Zq 0b − −
u̇2 /δe 46.05 −g sin(Θo ) -2.314a − −
c
Mu 0.0002066 − −
ẇ/δe 16.44
ẇ2 /δe 14.01 Mw 0b − −
θ/δe 34.27 Mq -0.5354 2.915 1.123
G-matrix
q/δf 19.12
ax/δf 11.47 Xde 0b − −
az/δf 49.06 Xdf 0b − −
u̇/δf 64.57 Xdt 0.1441a − −
Zde -1.746 11.08 5.083
ẇ/δf 25.14
Zdf -1.413 3.943 1.773
Jave 25.71 a
Zdt 0.01515 − −
Mde -0.1638 1.352 0.5923
Mdf 0.008682 2.245 0.9454
Mdt -0.00056a − −
Time delay
τe 0a − −
a
τf 0 − −
τT 0a − −
a Fixed parameter
b Eliminated parameter
c Constrained parameter

Table 3. Dynamic stability modes for 300 KIAS heavy load-


ing configuration.

Mode Damping (ζ) Frequency (ω) [rad/sec]


Spiral - 0.02441
Roll - 2.192
Dutch Roll 0.1606 2.384
Phugiod -0.0745 0.1081
Short Period 0.9863 0.5962

14 of 27

American Institute of Aeronautics and Astronautics


1. Time Domain Verification
With the completion of the model identification in the frequency domain, it was necessary to verify that
the model had good predictive capability and robustness to input variance in the time domain. The state-
space identified model was verified against flight test doublet data that was not used in the identification
process. Figure 17 shows flight data of the piloted stick δlon and the aerosurface command δe , which
was used to simulate the model. This comparison of the flight and model data shows that the model
captures the dynamics of the VISTA well, shown by the close agreement in the figures for relatively large
amplitudes (accelerations of 100 ft/sec2 ). Also, the model has RMS fit error cost JRMS = 1.8, where values

1
40

ax [ft/sec2 ]
/lon [inch]

0
!1

!2 -40

10
0

az [ft/sec2 ]
/e [deg]

-40
0

!5 -80

10 15
Flight Data Model
5 10
, [deg]

3 [deg]

0 5

-5 0

20
0
q [deg/sec]

10
w [ft/sec]

0
!50
-10
!100
-20
0 1 2 3 4 5 0 1 2 3 4 5

Time [sec] Time [sec]

Figure 17. Time domain verification results of pitch doublet (300 KIAS, 20 kft, heavy loading config-
uration).

of JRMS ≈ 1.0 to 2.0 reflect an acceptable level of accuracy.3 Time validation with doublets was performed
for all of the aerosurfaces of every point model. The RMS fit error costs were all below JRMS = 2, which
gives confidence in the identified models for both the longitudinal and lateral/directional axes.

IV. Stitched Model Background


A continuous, full flight envelope simulation model of the VISTA F-16 was developed to represent the aircraft
dynamics accurately over its entire flight and loading envelopes. This was accomplished using the model
stitching technique,2, 3, 5 which refers to the process of combining a collection of linear state-space models at
various flight conditions with trim data into a full envelope simulation model.

The stitched model architecture enables scaling of the forces and moments for changes in mass, moments
of inertia, and center of gravity (cg) location through the equations of motion, thus allowing for simulation
of changes in those parameters without explicitly including them as scheduling parameters. The resulting
model is accurate for typical aircraft flight dynamics for the entire maneuvering flight regime, except for

15 of 27

American Institute of Aeronautics and Astronautics


extreme conditions (e.g., stall or spin). A block diagram schematic of the stitched model is shown in Figure
18. One of the key features relevant to the stitched model is that the primary scheduling parameter is the

U Uf
Nonlinear
Control input • [identified
specific Variable Description
gravity forces
channels Trim values control derivatives]
(look-up table) (look-up table)
U Total longitudinal body axis velocity
m or msim Uf Filtered velocity
+ Δu BΔu
Pilot commands m or msim or
sim ∆u Control perturbations
m Total
+ + Forces
Nonlinear
∆x State perturbations
U Specific Aero + + (lb)
equations
Trim forces
+ Moments of motion M Mass and inertia matrix
(ft - lb)
Aircraft state Δx AΔx A Dimensional stability derivatives
+
Uf Low-pass U B Dimensional control derivatives
Aircraft state • [identified filter
Trim values stability derivatives]
m Aircraft mass
(look-up table) (look-up table)
I Aircraft inertia matrix
U Uf
Aircraft state

Figure 18. Model stitching block diagram schematic from Ref. 3.

total instantaneous x-body axis velocity state U or the true airspeed for typical angle of attack ranges. This
highlights the need for a fine grid of trim data around each point model to accurately represent the speed
stability derivatives (Xu , Zu , and Mu ). Additionally, the stitched model is able to simulate take-off and
landing through the inclusion of a simple landing gear and spoilers models, and different levels of wind and
turbulence.

Model stitching is accomplished by implementing lookup tables of the aircraft state trim values, control
input trim values, and stability and control derivatives based on the identified point models and trim data.
Perturbation aerodynamic and control forces and moments are determined from the perturbation states ∆x
and controls ∆u. The aerodynamic trim forces and moments are then summed to the perturbation values
to yield the total aerodynamic forces and moments acting on the aircraft in body axes.

V. Stitched Model from Flight Test Data


This section describes repurposing of the longitudinal static stability (LSS) data, scaling of the identified
models to a common loading configuration, culmination of the final stitched model data and validation.

A. Repurposing Longitudinal Static Stability Data


In the stitching architecture, a grid of trim data around each point model was generated using repurposed
longitudinal static stability data. For this study, each point model was trimmed at airspeeds of −20, −10,
0 +10, and +20 kts about its trim condition, which is at finer increments than was collected for during the
trim shot record. LSS data, discussed in Section III.A.1, was processed through the stitched model at a
constant altitude and straight-and-level (γ = 0), which produced a fine grid of constant altitude trim data or
repurposed LSS data. The process of repurposing LSS data to additional sets of constant altitude trim shot
data was repeated for each of the individual point models. The plot of the flight path angle γ0 in Figure 19
highlights that before repurposing the LSS data, the LSS data had small variations in flight path and almost
constant throttle, whereas the repurposed LSS data has varying throttle with constant altitude. Once all of
the LSS data were repurposed for each flight condition, the trim of the identified point models were updated.

16 of 27

American Institute of Aeronautics and Astronautics


LSS data ("h)
5 Repurposed LSS data (h_ = 0)

.0 [deg] /t0 [deg] /f0 [deg] /e0 [deg] ,0 [deg] 30 [deg]


4
3
5
4
3
2:4
2:3
2:2
2:05
2
1:95
43:4
43:2
43
2
0
!2
640 650 660 670 680 690 700 710 720
U [ft/sec]

Figure 19. LSS data and repurposed LSS data to constant altitude for 300 KIAS at 20 kft.

B. Baselining Models
After individual point models at each flight condition were identified, the models needed to be scaled to a
common loading configuration for implementation in the stitched model. The scaling technique demonstrated
in this section is explained in detail by Ref. 5, and performed in the same manner for the VISTA. The model
scaling technique was tested using the point models from the flight condition, 300 KIAS at 20 kft, at two
loading configuration: full and empty fuel tank.

1. Heavy and Light Loading Configuration Comparison


Figure 20 compares the primary roll rate frequency response for the VISTA flown with a full and empty fuel
tank.

p=/a
0
Magnitude [dB]

-10

-20

-30

-40
0
Light/Fwd .ight data
Phase [deg]

-90 Heavy/Aft .ight data


-180
-270
-360
1
Coherence [-]

0:5

0
10!1 100 101 102

Frequency [rad/sec]

Figure 20. Roll rate response to asymmetric flaperon for heavy and light loading configurations.

17 of 27

American Institute of Aeronautics and Astronautics


The constant shift in the high frequency magnitude response suggests that the fuel load variation causes
variations in the moment of inertia and changes the control derivative. As expected, the aircraft flown with
a full fuel tank has a lower magnitude at high-frequency compared to the empty fuel tank response, shown
in the roll rate response in Figure 20. This variation was captured in the identified point model stability
and control derivatives. Model scaling calculations for moments of inertia and cg values and ratios covered
in Ref. 5 were followed. For the scaling method, the identification results are used to determine the ratio
between the moments of inertia of the two configurations. For example, the control derivative for the full tank
was Lδa = −0.49 and for the empty tank was Lδa = −0.61. The ratio between the control derivatives was
used as an initial calculation for the ratio of the moments of inertia between the two loading configurations.
Lδa Light IxxHeavy
= ≈ 1.2 (12)
Lδa Heavy IxxLight
For the VISTA, a full fuel tank is associated with the heavy-weight/aft-cg configuration and the empty fuel
tank with the light-weight/fwd-cg configuration, shown in Table 4.

Table 4. Loading configuration comparison (300 KIAS, 20kft).

Light/Fwd Configuration Heavy/Aft Configuration


Total Fuel [lbs] 3,943 5,987
Total Weight [lbs] 25,343 27,387
Xcg [in] 322.4 324.9
Zcg [in] 91.4 93.2
Ixx [lbs-ft2] 280,813 346,621
Iyy [lbs-ft2] 2,302,200 2,360,079
Izz [lbs-ft2] 2,517,727 2,948,359
Ixz [lbs-ft2] 5,920 7,359

To test the ability of the stitched model to scale from one loading configuration to another the point model
identification of the light-weight/fwd-cg configuration was scaled to form the stitched model in the heavy-
weight/aft-cg loading configuration of the 300 KIAS flight condition. Figure 21 provided validation of the
scaling technique from the good agreement seen in the dynamic responses between the flight data and scaled
stitched model. All of the point models were scaled to the common loading configuration of heavy-weight/aft-
cg loading configuration for the final stitched model. When the point models were scaled to the common
loading configuration, the trim data at each flight condition was also updated to the common loading con-
figuration.

18 of 27

American Institute of Aeronautics and Astronautics


q=/e w=/f
0 30
Magnitude [dB]

Magnitude [dB]
-10 20
10
-20
0
-30
-10
-40 -20
-50 -30
0 0
Phase [deg]

Phase [deg]
-180 Light/Fwd Flight Data -180 Light/Fwd Flight Data
Light/Fwd ID Model Light/Fwd ID Model
Heavy/Aft Flight Data Heavy/Aft Flight Data
-360 -360
Heavy/Aft Stitched Model Heavy/Aft Stitched Model
-540 -540
1 1
Coherence [-]

Coherence [-]
0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

a) b)

p=/h r=/a
0 -20
Magnitude [dB]

Magnitude [dB]
-10 -30
-20 -40
-30 -50
-40 -60
-50 -70
0 180
Phase [deg]

Phase [deg]

-90 Light/Fwd Flight Data Light/Fwd Flight Data


0
-180 Light/Fwd ID Model Light/Fwd ID Model
Heavy/Aft Flight Data Heavy/Aft Flight Data
-180
-270 Heavy/Aft Stitched Model Heavy/Aft Stitched Model
-360 -360
1 1
Coherence [-]

Coherence [-]

0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

c) d)

p=/a r=/r
0 -20
Magnitude [dB]

Magnitude [dB]

-10 -30

-20 -40

-30 -50

-40 -60
0 0
Phase [deg]

Phase [deg]

-90 Light/Fwd Flight Data Light/Fwd Flight Data


-180
-180 Light/Fwd ID Model Light/Fwd ID Model
Heavy/Aft Flight Data Heavy/Aft Flight Data
-360
-270 Heavy/Aft Stitched Model Heavy/Aft Stitched Model
-360 -540
1 1
Coherence [-]

Coherence [-]

0:5 0:5

0 0
10!1 100 101 102 10!1 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

e) f)

Figure 21. a) q/δe b) w/δf c) p/δa d) r/δr e) p/δh f ) r/δa comparison between scaled stitched model,
flight identified model and flight data at 300 KIAS at 20 kft.

19 of 27

American Institute of Aeronautics and Astronautics


C. Final Stitched Model
The baselined models and trim data were collected and formed into lookup tables for the stitched model
integration. This section covers the last steps to form the complete stitched model, including smoothing of
stitched model derivatives and interpolation of trim data, and presents the final results.

1. Stability and Control Derivatives


Figures 22 and 23 show variation in stability and control derivatives with speed (x-body axis velocity U ). The
values of the stitched model derivatives shown in green are baselined to the nominal loading configuration,
and the identified parameter values are in red. During flight testing, the longitudinal axis maneuvers were
flown during the beginning of the flight, when the fuel tank was full.

u w q /e /f /T
0 0:08 1 1 0:05 0:17

!0:005 0:06 0 0:16


0 0
X

!0:01 0:04 !0:05 0:15

!0:015 0:02 !1 !1 !0:1 0:14

0 !0:4 2 0 0 0:018

!0:05 !0:6 !1
0 !2 0:016
Z

!0:1 !0:8 !2

!0:15 !1 !2 !3 !4 0:014
#10!4 #10!3
5 10 0 0 0:02 1
Heavy/Aft ID models
0 5 0 Baselined stitched model
M

!1 !0:2 0
!5 0 !0:02

!10 !5 !2 !0:4 !0:04 !1


300 600 900 300 600 900 300 600 900 300 600 900 300 600 900 300 600 900

U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec]

Figure 22. Longitudinal axis stitched model derivatives baselined to heavy-weight/aft-cg loading con-
figuration at 20 kft.

Whereas, the lateral/directional axis flight data was collected during the second half of the flights when the
aircraft was lighter. For this reason, the stitched model and identified parameters in Figure 23 show good
agreement in overall trends, but not exact values because the stitched model was baselined to a heavier
loading configuration than the light loading configurations of the lateral/directional identified models.

20 of 27

American Institute of Aeronautics and Astronautics


v p r /a /h /r
0:1 15 1 6 4 0

4
0:05 10 0 2 !0:5
Y

0 5 !1 0 0 !1

!0:04 !1 2:5 0 0 0:4

!2 2 !0:5
!0:06 !0:5 0:2
L

!3 1:5 !1

!0:08 !4 1 !1:5 !1 0
#10!3 #10!3
10 0:05 !0:2 0 0 0

!0:4
5 0 !2 !0:05 !0:1
N

!0:6 Light/Fwd ID models


Baselined stitched model
0 !0:05 !0:8 !4 !0:1 !0:2
Heavy/Aft ID model
300 600 900 300 600 900 300 600 900 300 600 900 300 600 900 300 600 900

U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec] U [ft/sec]

Figure 23. Lateral/directional axis stitched model derivatives baselined to heavy-weight/aft-cg loading
configuration at 20 kft.

2. Final Trim Data Set


The five sets of repurposed LSS data from the baselined point models were combined for additional trim
data. Figure 24 shows the close agreement between the trim data sets of baselined and repurposed LSS data
(stitched model trim data) compared to the trim shot data collected at constant altitude. The LSS data
repurposed to trim data were combined by fitting each trim value with least squares fits over the various
airspeeds. Special attention should be given to the bottom plot in Figure 24, where the flight path from
the trim shot record was not exactly flat, whereas the repurposed LSS data by design has a zero flight path
angle (γ0 = 0). The fits enabled extrapolation and interpolation of the trim values to produce a finer trim
data set for the stitched model.

21 of 27

American Institute of Aeronautics and Astronautics


Trim shot data
20 Repurposed LSS data

.0 [deg] /t0 [deg] /f0 [deg] /e0 [deg] ,0 [deg] 30 [deg]


10
0
20
10
0
3
2
1
2:2
2
1:8
60
40
20
2
0
!2
400 500 600 700 800 900
U [ft/sec]

Figure 24. Repurposed LSS trim data sets compared to trim shot data collected at a constant altitude.

D. Validation of Stitched Model


Validation of the stitched model was performed at several different flight conditions. The flight data used
in this validation were recorded at a different loading configurations (empty fuel tank). Therefore, the
results presented validate the ability of the stitched model to trim at a simulated loading configuration,
which requires extrapolation. To generate the stitched model validation results, the model was trimmed at
the same flight condition and loading configuration as the flight data. The aerosurface deflections recorded
during piloted doublets were used to simulate the stitched model. Figure 25 shows the pitch axis doublet
response of the stitched model, scaled to the light-weight/fwd-cg loading configuration at the 300 KIAS,
20kft flight condition, which closely agrees with the flight data. This validation was performed for each of
the aerosurfaces, resulting in low RMS fit error costs of JRMS < 2. This ensured that the stitched model
can be extrapolated to different loading configurations and accurately capture the dynamic characteristics
of the VISTA.

22 of 27

American Institute of Aeronautics and Astronautics


1
40 Flight Data Stitched Model

ax [ft/sec2 ]
/lon [inch]
0

0
!1

!2 -40

10
0

az [ft/sec2 ]
/e [deg]

5
-40
0
-80
!5

15 15
10 10
, [deg]

3 [deg]
5 5
0 0
-5 !5

20
0
q [deg/sec]

10
0 w [ft/sec] !50
-10
!100
-20
0 2 4 6 8 0 2 4 6 8

Time [sec] Time [sec]

Figure 25. Stitched model pitch doublet validation (300 KIAS, 20 kft, empty loading configuration).

VI. Stitched Model Trends/Analysis


From the final stitched model, trends over the complete full flight envelope can be analyzed and insight
into the characteristics of the aircraft can be drawn. A point of interest for the VISTA is how its stability
changes with speed. Interpolation of the stitched model to various airspeeds is shown in Figures 26−28.
The stability of the aircraft is depicted in Figures 29 and 30 where the longitudinal and lateral/directional
axes eigenvalues at the various airspeeds are presented. As shown in Figure 26, as speed is increased, the
on-axis pitch response has an increase in high-frequency pitch response magnitude and short-period natural
frequency. The short period mode changes from a heavily damped stable mode at 200 KIAS (two stable poles
at −0.15 and −0.59 rad/sec) to an unstable mode at 420 KIAS (stable pole at −4.1 rad/sec and unstable
pole at 1.2 rad/sec) as speed is increased; see Figure 29. Notably, the phugoid mode is only visible in Fig-
ure 26 for the lowest speed, just above 0.11 rad/sec. As the speed increases, the phugoid frequency decreases.

Variation in the roll axis with airspeed is shown in Figure 27. As expected, the roll mode/Dutch roll
mode dipole increases in frequency with increased airspeed.3 Figure 30 shows that the lightly damped Dutch
roll mode becomes less damped with increased speed. From the shift in the yaw responses to the right
(Figure 28), one can deduce that the frequency of the Dutch roll mode frequency increases with speed, which
is also seen in Figure 30 by the increasing natural frequency of the Dutch roll mode.

23 of 27

American Institute of Aeronautics and Astronautics


q=/e p=/a
0 0

!10
!10
Magnitude [dB]

Magnitude [dB]
!20
!20
!30
!30
!40

!50 !40
Increasing speed Increasing speed
!60 !50

200 KIAS 200 KIAS


-90 250 KIAS -90 250 KIAS
Phase [deg]

Phase [deg]
300 KIAS 300 KIAS
360 KIAS 360 KIAS
420 KIAS 420 KIAS
-270 -270

-450 !1 -450 !1
10 100 101 102 10 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

Figure 26. Pitch rate response identification results Figure 27. Roll rate response identification results
at varying speed. at varying speed.

r=/r
!10

!20
Magnitude [dB]

!30

!40

!50

!60
Increasing speed
!70

200 KIAS
-90 250 KIAS
Phase [deg]

300 KIAS
360 KIAS
420 KIAS
-270

-450 !1
10 100 101 102

Frequency [rad/sec]

Figure 28. Yaw rate response identification results


at varying speed.

4 4

3 200 KIAS 3 200 KIAS


250 KIAS Phugoid Mode 250 KIAS
2 300 KIAS 2 300 KIAS
Imaginary Axis [sec!1 ]

Imaginary Axis [sec!1 ]

360 KIAS 360 KIAS


1 420 KIAS 1 420 KIAS
Spiral Mode
0 0

!1 !1 Roll Mode w/ Increasing Speed

!2 Short Period Mode w/ !2


Increasing Speed
!3 !3
Dutch Roll Mode w/ Increasing Speed
!4 !4
!5 !4 !3 !2 !1 0 1 2 !3:5 !3 !2:5 !2 !1:5 !1 !0:5 0
Real Axis [sec!1 ] Real Axis [sec!1 ]

Figure 29. Longitudinal axis eigenvalues. Figure 30. Lateral/directional axis eigenvalues.

24 of 27

American Institute of Aeronautics and Astronautics


When flight conditions are below Mach number M = 0.5, most of the variations in fixed-wing derivatives
are due to altitude and airspeed and therefore the derivatives are fairly constant below M = 0.5.3 Similar
to analysis of the fixed-wing Cessna Citation in Ref. 3, the nondimensional longitudinal derivatives vary
significantly with Mach number above M = 0.5. By analyzing the nondimensional derivatives, taking away
the dependency on airspeed and density, one can see how the dynamics of the aircraft change independent
of airspeed and density.3 As the VISTA increases in Mach number, the nondimensional derivatives associ-
ated with longitudinal stability change dramatically. Figure 31 shows that as Mach number increases, the
nondimensional static stability pitching-moment derivative CMα becomes increasingly positive (unstable)
and the nondimensional derivative CMu becomes increasingly negative, which also contributes to the aircraft
becoming unstable.12 This rapid increase in magnitude and change in sign of derivative CMu is referred to
as Mach “tuck”.12

Figure 31. Nondimensional pitching-moment derivatives.

Figure 31 suggests that at low Mach number the nondimensional derivatives change very little, but dramatic
changes occur above Mach 0.6. Therefore, the point model identification speeds could be arranged further
apart over the low Mach number range and more finely at higher speeds, corroborating conclusions made in
References 2 and 3.

Next, the ability of the stitched model to extrapolate for different weight and cg was demonstrated. Here,
the stitched model was used to predict the pitch rate frequency response for the full range of loading con-
figurations (from full to empty fuel tank). Figures 32 and 33 show a mixture of heavy and light identified
models and stitched model extrapolated for various loading configurations.

q=/e w=/f
0 60
28400 lbs Stitched model 28400 lbs Stitched model
-10 27387 lbs Heavy ID model 40 27387 lbs Heavy ID model
Magnitude [dB]

Magnitude [dB]

26925 lbs Stitched model 26925 lbs Stitched model


-20 25343 lbs Light ID model 20 25343 lbs Light ID model
22500 lbs Stitched model 22500 lbs Stitched model
-30 0

-40 Increasing weight !20 Increasing weight

-50 !40
0 0

-90 -90
Phase [deg]

Phase [deg]

-180 -180

-270 -270

-360 -360

-450 !1 0 1 2
-450 !1
10 10 10 10 10 100 101 102

Frequency [rad/sec] Frequency [rad/sec]

Figure 32. Stitched model pitch rate responses for Figure 33. Stitched model vertical velocity re-
different fuel levels for 300 KIAS at 20 kft. sponses for different fuel levels for 300 KIAS at
20 kft.

25 of 27

American Institute of Aeronautics and Astronautics


The effect of weight variation on the longitudinal response is shown in Figures 32 and 33. The lighter
weight configurations have a more forward cg, which shifts the short period frequency and decreases the
response magnitude (at around 0.2 − 0.9 rad/sec). These trends are well characterized by the stitched model
extrapolations.

VII. Conclusion
This paper covered the development of a full flight envelope simulation model of the NF-16D VISTA from
flight data. The stitched model was developed from discrete point models identified with the control system
engaged. Additional conditioning of the data was performed to deal with highly correlated aerosurface inputs.
A scaling method was implemented to baseline the models to a common loading configuration, making it
possible to interpolate and extrapolate the final stitched model over the full flight envelope for various flight
conditions and loading configurations.
1. Due to the highly correlated aerosurface, decorrelation of aerosurface inputs using Joint Input/Output
method was a necessary and effective method to extract frequency responses for system identification.
These results show it is possible to identify accurate bare-airframe models from closed-loop data with
high levels of correlation.
2. Implementation of the frequency response system identification method produced accurate point models
at each flight condition, which reached up to approximately Mach 0.9. The identify reliable stability
and control derivatives are very reliable as evident in the low Cramér-Rao bounds and low time domain
RMS fit cost of the models.
3. Longitudinal static stability data provided trim gradients of the states and controls for the determi-
nation of speed derivatives, Xu , Zu and Mu . Repurposed LSS data also provided a richer set of trim
data as needed for the stitched model.
4. The final stitched model was formed with stability and control derivatives and trim data baselined to
a common loading configuration. Validation results in the time and frequency domain show excellent
agreement between the stitched model and flight data. These results show the ability of stitched model
to simulate the aircraft over its full flight envelope, capturing the aircraft’s dynamics from low speed
up to high subsonic speeds with five identified point models and trim data.

Acknowledgments
The authors would like to thank Jason Kirkpatrick and Jay Kemper (Calspan Corporation) for their hard
work on the simulator and flight tests, as well the test pilots Roger Roscoe Tanner, Evan Ivan Thomas and
Evil Bill Gray for flying excellent frequency sweeps.

References
1 Hehs, E., “F-16 Designer Harry Hillaker,” Code One Online Articles, Lockheed Martin Aeronautics Company, April 1991.
2 Tobias, E. L. and Tischler, M. B., “A Model Stitching Architecture for Continuous Full Flight-Envelope Simulation of
Fixed-Wing Aircraft and Rotorcraft from Discrete-Point Linear Models,” U.S. Army AMRDEC Special Report RDMR-AF-16-
01, April 2016.
3 Tischler, M. B. and Remple, R. K., Aircraft and Rotorcraft System Identification: Engineering Methods and Flight Test

Examples Second Edition, AIAA, 2012.


4 Raol, J. R., Girija, G., and Singh, J., Modelling and Parameter Estimation of Dynamic Systems, The Institution of

Engineering and Technology, 2004.


5 Berger, T., Tischler, M. B., Hagerott, S. G., and Cotting, M. C., “Development and Validation of a Flight Identified

Business Jet Simulation Model Using a Stitching Architecture,” presented at AIAA Modeling and Simulation Technologies
Conference, January 2017.
6 Jategaonkar, R. V., Unstable Aircraft Identification, Flight Vehicle System Identification: A Time-Domain Methodology,

AIAA, 2015.
7 Hersey, S., Celi, R., Juhasz, O., Tischler, M. B., Rand, O., and Khromov, V., “State-Space Inflow Model Identification

and Flight Dynamics Coupling for an Advanced Coaxial Rotorcraft Configuration,” presented at the American Helicopter
Society 73rd Annual Forum, Fort Worth, TX, May 2017.
8 Forssell, U. and Ljung, L., “Closed-Loop Identification Revisited,” Automatica, Vol. 35, Issue 7, July 1999.

26 of 27

American Institute of Aeronautics and Astronautics


9 Akaike, H., “Some Problems in the Application of the Cross-Spectral Method,” Spectral Analysis of Time Series, edited

by B. Harris, John Wiley, New York, 1967.


10 Anon., “Handling Qualities Requirements for Military Rotorcraft,” Aeronautical Design Standard-33 (ADS-33E-PRF),

US Army Aviation and Missile Command, March 2000.


11 Stevens, B. and Lewis, F., Aircraft Control and Simulation, Wiley and Sons, 1992.
12 McRuer, D., Ashkenas, I., and Graham, D., Aircraft Dynamics and Automatic Control, Princeton University Press, 1990.

27 of 27

American Institute of Aeronautics and Astronautics

View publication stats

You might also like