Nomenclature
-
$v$
-
flight velocity (m/s)
-
$\gamma $
-
flight path angle (rad)
-
$h$
-
flight altitude (m)
-
$\alpha $
-
angle-of-attack (rad)
-
$q$
-
fitch rate (rad/s)
-
${\delta _e}$
-
elevator deflection angle (rad)
-
$\beta $
-
throttle setting
-
$T$
-
thrust (N)
-
$L$
-
lift (N)
-
$D$
-
drag (N)
-
$m$
-
mass (kg)
-
$c$
-
mean aerodynamic chord (m)
-
${M_y}$
-
pitching moment (
${\rm{kg}} \cdot {{\rm{m}}^2}$ )
-
${R_e}$
-
radius of the Earth (m)
-
$r$
-
radial distance from Earth’s centre (m)
-
$\rho $
-
density of air (kg/m
${{\rm{\;}}^3}$ )
-
$a$
-
speed of sound (m/s)
-
$M$
-
mach number
-
${I_y}$
-
moment of inertia (
${\rm{kg}} \cdot {{\rm{m}}^2}$ )
-
$\mu $
-
gravitational constant (
${{\rm{m}}^3}/{s^2}$ )
-
${C_L}$
-
lift coefficient
-
${C_D}$
-
drag coefficient
-
${C_T}$
-
thrust coefficient
-
${C_{M\alpha }}$
-
derivative of pitching moment coefficient with respect to angle-of-attack
-
${C_{Mq}}$
-
derivative of pitching moment coefficient with respect to pitch rate
-
${C_{M{\delta _e}}}$
-
derivative of pitching moment coefficient with respect to elevator
-
${\rm{ADRC}}$
-
active disturbance rejection control
-
${\rm{LPV}}$
-
linear parameter varying
-
${\rm{ITAE}}$
-
integral of time absolute error
-
${\rm{LQR}}$
-
linear quadratic regulator
1.0 Introduction
Wide-envelope flight vehicles operating in near-space have attracted significant attention owing to their high flight and response speeds and outstanding manoeuverability [Reference Shahzad and Hu1,Reference Mimouni, Araar, Oudda and Haddad2]. These types of aircraft must manoeuver over a wide range of altitudes and speeds to meet real-time mission demands [Reference Wang and Xu3,Reference Suicmez and Kutay4], making the flight control a challenge [Reference Feng, Sun, Wu, Wang and Xi5,Reference Wang and Xia6]. In addition, the near-space environment is complex [Reference Ni and Fang7], changeable and uncertain; thus, control-oriented models have large perturbation ranges and multiple disturbance sources with wide variation range [Reference Qu, Wang, Liu, Fu and Bai8]. However, this results in significant variations in their dynamic characteristics [Reference Wu, Zhou and Wang9], making it difficult for a single controller to cover all operating conditions [Reference Wang, Fan, Zhang, Fan and Yan10–Reference Li, Liu, Gong, Zhao and Yuan12]. Therefore, there is an urgent need for a wide-envelope automatic gain-scheduling strategy for flight vehicles with model uncertainty [Reference Xu and Shi13,Reference Xu, Shou, Shi and Yan14].
To overcome unknown disturbances and uncertainties during flight, the active disturbance rejection control (ADRC) proposed by Han is widely used in aircraft attitude control [Reference Han15]. ADRC employs an extended state observer (ESO) to estimate and compensate for modeling errors and nonlinear uncertainties without relying on specific mathematical models [Reference Sun, Wang and Chen16]. Gao simplified the initial nonlinear ADRC into linear ADRC (LADRC) to facilitate its application in practice [Reference Gao17,Reference Huang and Xue18]. Although LADRC gain scheduling within a large flight envelope is intuitive and straightforward, designing the parameter set is difficult. Therefore, there is a need for an automatic gain-scheduling strategy for wide-envelope aircraft within the active disturbance rejection control framework.
For traditional gain-scheduling control, a linear controller is designed for each operating point; thus, it is difficult to ensure global control performance. In addition, it highly depends on the selected operating points [Reference Xie, Wei, Gu and Shi19,Reference Ducard and Allenspach20]. The linear parameter-varying (LPV) technique embeds dynamic nonlinearities in the varying parameters, explicitly accounting for the variations in real-time parameters. Recently, the LPV method has been widely used for gain-scheduling control of flight vehicles owing to its effectiveness in managing variations. Caigny [Reference Caigny, Camino, Oliveira, Peres and Swevers21] proposed a gain-scheduling controller based on a parametric-dependent Lyapunov function matrix. This method was successfully applied to an aircraft LPV model by solving parametric-dependent linear matrix inequalities (LMI). Kazemi [Reference Kazemi and Tarighi22] developed a polyhedral LPV for quadrotor aircraft with uncertainties to achieve robust attitude control. Lucas [Reference Souza, Peixoto and Palhares23] proposed a sufficient stability condition for time-delayed LPV systems based on the Lyapunov–Krasovskii function. Jiang [Reference Jiang, Zheng, Sun and Wang24,Reference Jiang, Wu, Wang and Wang25] employed the switching polyhedral linear parameter-varying (SPLPV) framework for gain scheduling of the output feedback control to ensure stability and robustness throughout the flight envelope. Vinco [Reference Vinco, Sename and Strub26] proposed an LPV control design for guided projectiles. They used an LPV polytopic system to ensure stability and performance across the flight envelope while reducing the computational complexity of the controller synthesis. These studies have shown that converting the nonlinear model into a polytopic LPV model can effectively address the practical challenge of controller design for morphing aircraft, thereby offering significant engineering application values. The methods developed in previous studies focused on Hurwitz’s stability with respect to parametric uncertainty in specific convex geometrical regions.
The Guardian map (GM) defines a scalar-valued function of parameterised matrices with a specific distribution of eigenvalues or roots known as generalised stability. Saydy [Reference Saydy, Tits and Abed27] developed the Guardian map as a unifying tool for studying generalised stability. It can quantitatively analyse the range of scheduling variables for LPV stability and has been extensively employed in control designs owing to its concise algebraic representation of closed-loop performance defined by generalised stability, such as flying quality (FQ). Saussie [Reference Saussie, Ouassima and Saydy28] verified the effectiveness of the GM theory for robust generalised stability problems using a second-order linear model. Xiao [Reference Xiao, Liu, Liu and Lu29–Reference Chen, Chen and Liu31] employed this theory to design the control law of large-envelope aircraft and achieved adaptive tuning of the control parameters. Cao [Reference Cao, Wan, He and Lu32] proposed a multimodel switching predictive control strategy based on GM to ensure successful perching maneuvers and smooth switching.
Herein, we propose a flight vehicle ADRC gain-scheduling strategy based on GM. The gap metric theory [Reference French33] was adopted to more accurately describe the discrepancy between linear systems, making the LPV model more precise and reasonable than the traditional Jacobian linearisation method. The GM was employed to calculate the initial controller gains at a boundary point. Through iterations, a controller set covering the entire flight envelope can be obtained automatically, and a controller gain-scheduling strategy can be achieved. The primary contributions of this paper are as follows:
-
(1) The gap metric-based nominal point selection (NPS) algorithm is used to divide the envelope and determine the linearised operating points, ensuring accurate LPV modeling while reducing computational complexity and enhancing model reliability.
-
(2) The GM is employed in ADRC gain scheduling, which can automatically extend the performance of an initial controller to the entire two-dimensional (2D) envelope with enhanced robustness, thereby automating the design process without manual intervention.
The remainder of this paper is organised as follows. Section 2 presents the nonlinear model for wide-envelope flight vehicles. Section 3 presents the LPV model using the gap metric. In section 4, the gain-scheduling algorithm for ADRC based on GM is proposed. Section 5 presents the simulation results, and Section 6 concludes the paper.
2.0 Longitudinal nonlinear modeling for flight vehicle
The winged-cone vehicle model, which was used in the NASA wind experiment for aerodynamic testing [Reference Shaughnessy, Pinckney and McMinn34], was selected as the subject of this study. Figure 1 shows the vehicle and its control surfaces (elevon and rudder). The elevon is a combination of the elevator and aileron; thus, it performs both pitch and roll control functions. Table 1 lists its parameters. The equations of the longitudinal motion are expressed as follows [Reference Marrison and Stengel35]

with

where
$v$
,
$\gamma $
,
$h$
,
$\alpha $
and
$q$
are the flight speed, flight path angle, flight altitude, angle-of-attack and pitch rate of the aircraft, respectively;
$L$
,
$D$
,
$T$
and
${M_y}$
represent the lift, drag, thrust and pitching moments, respectively;
$m$
,
${I_y}$
,
$\mu $
and
${R_e}$
denote the vehicle mass, longitudinal moment of inertia, gravitational constant and earth radius, respectively.
${C_L}$
,
${C_D}$
and
${C_T}$
are the coefficients of the lift, drag, and thrust, respectively, and
${C_{M\alpha }}$
,
${C_{Mq}}$
and
${C_{M{\delta _e}}}$
are the derivatives of the pitching moment with respect to the angle-of-attack, pitch rate and elevator deflection, respectively. The variables in Equation (2) are as follows:

Table 1. The parameters in the model


Figure 1. Configuration geometry.
The dynamic equation of the elevator actuator is expressed as follows:

where
${\delta _c}$
is the elevator command,
${\delta _e}$
the elevator deflection angle,
${\omega _a}$
the bandwidth and
${\xi _a} = 0.707$
the damping ratio.
3.0 Gap metric-based aircraft lpv modeling
3.1 Gap metric
Let
$K$
be a linear operator in Hilbert space
$H$
[Reference French33]:

$H \times H$
is the set of all pairs
$u,v$
;
$u$
and
$v$
are in
$H$
.
$H \times H$
is also a Hilbert space with the inner product derived from
$H$
. The domain of
$K$
,
$D\left( K \right)$
, is the set of all
$u$
in
$H$
with the property that
$Ku$
is in
$H$
. The graph
$G\left( K \right)$
of
$K$
is the set of all pairs
$u,Ku$
with
$u$
in
$D\left( k \right)$
and is defined as
$G\left( K \right) = \{ \left( {u,Ku} \right) \in H \times H|u \in D\left( K \right)\} $
.
Because
$K$
is linear,
$G\left( k \right)$
is a subspace of the product Hilbert space
$H \times H$
.
$K$
is considered closed if its graph
$\left( K \right)$
is a closed subspace of
$H \times H$
.
Definition 1.
The gap between two closed operators
${K_1}$
and
${K_2}$
in Hilbert space
$H$
is defined as the gap between their graphs considered closed subspaces:

where
${\delta _{12}}\left( {G\left( {{K_1}} \right),G\left( {{K_2}} \right)} \right) = \mathop {{\rm{sup}}}\limits_{u \in D\left( {{K_1}} \right),u \ne 0} \mathop {{\rm{inf}}}\limits_{v \in D\left( {{K_2}} \right)} \frac{{u - {v^2} + {K_1}u - {K_2}{v^2}}}{{\sqrt {{u^2} + {K_1}{u^2}} }}.$
${\delta _{21}}\left( {G\left( {{K_2}} \right),G\left( {{K_1}} \right)} \right)$
is defined similarly.
Example 1:
${K_1} = \left[ {{\rm{cos}}\theta, - {\rm{sin}}\theta ;\,{\rm{sin}}\theta, {\rm{cos}}\theta } \right]$
is a rotation by
$\theta $
radians.
${K_2} = \left[ {\lambda, 0;\,0,\lambda } \right]$
is a scaling by a factor of
$\lambda $
.
$u$
and
$v$
are the unit vectors on the unit circle
${\mathbb{R}^2}$
. Figure 2 shows a gap
$\delta \left( {{K_1},{K_2}} \right)$
.

Figure 2. Geometric representation of operators
${K_1}$
and
${K_2}$
with gap.
We consider linear systems as mappings in Hilbert space. For a multiinput multioutput linear system, which is represented by state matrices
$A$
and
$B$
and input matrices
$C$
and
$D$
, the transfer function is expressed as
$P\left( s \right) = C{\left( {sI - A} \right)^{ - 1}}B + D$
, where
$I$
is the identity matrix. Notably,
$P\left( s \right)$
is a linear operation in Hilbert space.
Definition 2.
If
${P_1}$
and
${P_2}$
are the transfer functions of two systems, the gap between the two systems can be expressed by the operator gap as follows:

where

and
$\left( {{M_1},{N_1}} \right)$
,
$\left( {{M_2},{N_2}} \right)$
represent the left prime decomposition of the linear transfer functions
${P_1}$
and
${P_2}$
, respectively.
$Q$
is a Hilbert matrix.
The properties of the gap metric include the following (refer to Ref. (Reference French33,Reference Tao, Li and Li36) for more details):
-
(1) The gap metric is a measure of the distance between two linear systems.
-
(2) The gap metric lies between 0 and 1,
$0 \le {\delta _{gap}}\left( {{P_1},{P_2}} \right) \le 1$ . The smaller the gap between the two systems, the closer their dynamic characteristics are.
3.2 Gap metric-based NPS algorithm for LPV modeling
The GM was used to analyse the generalised stability of parametric matrix families, which requires state space models. Therefore, it is crucial to develop an efficient LPV to cover nonlinear dynamics.
A nonlinear model in the LPV form can be represented using three methods: Jacobian linearisation, state transformations and function substitution [Reference Cavanini, Ippoliti and Camacho37]. Among them, Jacobian linearisation is widely used. However, it relies on a set of linearised models corresponding to several equilibrium points, which must cover the entire flight envelope to ensure good fidelity with the original nonlinear model. Therefore, to avoid choosing similar equilibrium points and reduce unnecessary computational complexity, the gap metric is introduced to measure the similarity among linear systems based on Jacobian linearisation, which is called the NPS algorithm. The NPS algorithm employs gap metric analysis to identify optimal nominal points within the sub-envelopes, ensuring accurate LPV modeling of nonlinear systems by minimising computational complexity and enhancing model fidelity.
First, the flight envelope is divided into subenvelopes according to the gap metric value. Then, a point in each subenvelope with the smallest difference in model characteristics with other state points is selected. This point is called the nominal point. Finally, the LPV model is obtained by fitting the linear model at nominal points using the following steps:
Suppose the large flight envelope is denoted as
$\phi $
, and its range is
$\left[ {{M_{{\rm{min}}}},{M_{{\rm{max}}}}} \right] \times \left[ {{h_{{\rm{min}}}},{h_{{\rm{max}}}}} \right]$
.
Step 1. In the flight envelope, the average grid points are divided at certain intervals of
${\Delta}M$
and
${\Delta}h$
, and a total of
${N_{Ma}} \times {N_h}$
state points are obtained.

Step 2. The corresponding LTI set is obtained using the Taylor series expansion method at the grid points, which can be denoted as
${P_{ij}},i = 1,2, \ldots {N_{Ma}}$
,
$j = 1,2, \ldots {N_h}$
.
Step 3. Calculate the gap measure between adjacent points using the expression
${\delta _{gap}}\left( {{P_{i,j}},{P_{i,j + 1}}} \right)$
or
${\delta _{gap}}\left( {{P_{i,j}},{P_{i + 1,j}}} \right)$
.
Step 4. Divide the envolpe into subenvelopes.
The altitude interval is divided only at
${M_{{\rm{min}}}}$
to avoid divergence between the Mach and altitude directions. When the cumulative value of the gap metric reaches a threshold
$\tau $
, expressed as

the corresponding interval numbers can be considered the altitude boundaries of the subenvelope to divide the entire envelope, denoted by the set
${h_B}$
. Here,
${h_B} = \left\{ {{h_{B1}},{h_{B2}}, \ldots, {h_{Bn}}} \right\}$
, where each
${h_{Bi}}$
is a boundary value in the altitude direction.
${i_0}$
is the starting index for the summation, and
$k$
is the number of intervals required for the cumulative gap metric to reach the threshold
$\tau $
. After reaching the threshold
$\tau $
, the next accumulation starts at
${i_0} = k + 1$
.
The speed interval is then divided at the altitude boundaries in the set
${h_B}$
. Similarly, if

the corresponding interval numbers can be considered the Mach boundaries. Here,
${j_0}$
is the starting index of the summation. After reaching the threshold
$\tau $
, the next accumulation starts at
$k + 1$
. For each altitude boundary, the Mach values are divided into multiple intervals. They can be considered as the boundaries under the corresponding altitudes, denoted by the sets
${M_{B1}}$
,…,
${M_{Bn}}$
. The subintervals are then numbered in ascending order of altitude and Mach. Thus,
$\phi $
can be divided into
$L$
subenvelopes based on these boundaries, denoted as
${\phi _l},l = 1,2, \ldots, L$
.
This process ensures a systematic division of the envelope
$\phi $
into subenvelopes with multiple Mach intervals under each altitude boundary, thereby providing detailed and flexible segmentation, as shown in Fig. 3.

Figure 3. Division of the envelope into subenvelopes.
Step 5. Select nominal points.
The points within the subenvelope
${\phi _l}$
that is least different from the other point linear models, called the nominal point
${E_l}$
, are calculated. Notably, the nominal points must satisfy Equation (11)

where
${n_M}$
and
${n_h}$
are the numbers of state points in
${\phi _l}$
along the velocity and altitude dimensions, respectively, and
$a$
and
$b$
index all state points in the
${\phi _l}$
.
${E_a}$
is set to the nominal point being considered, and
${E_b}$
represents all other state points within the subenvelope, except the nominal point.
This ensures that the selected nominal point
${E_l}$
best represents the subenvelope
${\phi _l}$
, resulting in an accurate LPV model.
Step 6. Obtain the LPV model of the aircraft by fitting the LTI system corresponding to the nominal points.

Figure 4. Variation of gap metrics with Mach numbers at various altitudes.

Figure 5. Variation of gap metrics with altitudes at various Mach numbers.
In this study, the NPS algorithm was used to derive the LPV model. This is achieved by linearising a set of LTI systems covering the region of interest, resulting in an LPV model that locally approximates the original nonlinear system around specific equilibrium points.
3.3 LPV model
The large flight envelope considered herein is
$h = \left[ {24,34} \right]$
km,
${\rm{\;\;}}M = \left[ {5,10} \right]$
, which corresponds to a classic hypersonic flight envelope [Reference Gao, Chen, Sun, Huang, Wang and Chen38]. Using the algorithm proposed in section 3.2, we analysed the trend of the gap metrics under constant speed or altitude conditions. Figures 4 and 5 show the variation in the gap metrics with altitude and speed, respectively. As shown in Fig. 4, increasing the Mach number generally reduces the gap metric, indicating that the difference in model characteristics due to speed decreases with increasing speed, and this difference is more obvious at low altitudes. At different speeds, the trend of the clearance measures with altitude was similar, as shown in Fig. 5. With increasing altitude, the differences in the model characteristics due to the change in altitude increase, especially at lower Mach numbers. This indicates that the impact of altitude on the model characteristics is more pronounced at lower speeds.
Here, we consider
${\Delta}M = 0.5$
,
${\Delta}h = 1$
, and
$\tau = 0.37$
; thus,
$N_{Ma} = N_{h} = 11$
, which corresponds to 121 grid points. According to gap metric theory, the flight envelope is divided into 16 subenvelopes, and the nominal points of each subenvelope are calculated, as shown in Fig. 6. The blue and red points in the figure indicate the grid and nominal points, respectively. Taking the second sub-envelope as an example, the gap metrics between each state point and other state points are calculated, and the mean value is recorded (Fig. 7). Here, point
${E_2}$
(
$h = 25{\rm{km}}$
,
$M = 6.5$
) has the smallest gap metric compared to the other state points, indicating that the model characteristics at point
${E_2}$
have the least variation from the rest of the points within the subenvelope. Therefore, the point with the minimal gap metric value
${E_2}$
can be selected as the nominal point.

Figure 6. Flight envelope division and schematic of the nominal points.

Figure 7. Gap metric mean value of the second subenvelope.
To simplify the range of values, the time-varying parameters
$M$
,
$h$
are normalised as follows:

where
${\delta _M} \in \left[ { - 1,1} \right]$
,
${\delta _h} \in \left[ { - 1,1} \right]$
.
Several LTI systems can be obtained by Jacobi linearisation at 16 equilibrium points in the flight envelope, and the LPV model is established by fitting them with
${\delta _M}$
and
${\delta _h}$
as the scheduling parameters:

with
$X = {\left[ {v,\gamma, h,\alpha, q} \right]^T}$
,
$U = {\left[ {\sigma, {\delta _e}} \right]^T}$
,

where
$\sigma $
is the fuel equivalent ratio,
${\delta _e}$
is the elevator angle, and
${a_{11}}$
,
${a_{13}}$
,
${b_{11}}$
, etc., are functions of
${\delta _M}$
and
${\delta _h}$
(see Appendix).
4.0 Adaptive gain scheduling of adrc with guardian map
4.1 Design of active disturbance rejection controller
Figure 8 shows the gain-scheduling framework, which comprises three key components: LADRC, LPV modeling based on gap metrics and gain scheduling using the GM. LADRC is central to vehicle attitude control, which is the core control method in this study. LPV models, which vary with vehicle altitude and speed, provide the foundation for calculating a stable region through the GM. The GM then dictates how control gains should be scheduled according to changes in altitude and speed, ensuring that the ADRC can meet the performance requirements of flight control.

Figure 8. Framework of the gain-scheduling control.
The primary objective of this study is to control the attitude (angle-of-attack) of the aircraft. Therefore, an LPV model is extracted for the short-period mode, with only the elevator deflection angle
${\delta _e}$
as input and the angle-of-attack
$\alpha $
and pitch rate
$q$
as state variables. Considering the control of the angle-of-attack, we designed an active disturbance rejection controller. According to the nonlinear model, the differential equation that describes the dynamics of
$\alpha $
can be expressed as follows:

Let
${b_0} = 0.5\rho {v^2}sc \times 0.0292/{I_y}$
,
${x_1} = \alpha $
,
${x_2} = \dot \alpha $
. According to the total disturbance concept of active disturbance rejection control, let
$f = 0.5\rho {v^2}sc\left( {{C_{M\alpha }} + {C_{Mq}}} \right)/{I_y} - \ddot \gamma $
be the total disturbance, defined as the new state
${x_3}$
.
Equation (14) can then be extended to the state-space form as follows:

Based on the idea of online disturbance compensation, the following extended state observer is designed to estimate the total disturbance in real time:

where
${z_i}\left( {i = 1,2,3} \right)$
is the estimate of
${x_i}\left( {i = 1,2,3} \right)$
, and
$L = {\left[ {{l_1},{l_2},{l_3}} \right]^\intercal}$
is the observer gain. To simplify the tuning process, the observer gains can be parameterised as
$L = {\left[ {{l_1},{l_2},{l_3}} \right]^\intercal} = {\left[ {3{\omega _o},3{\omega _o}^2,{\omega _o}^3} \right]^\intercal}$
[Reference Gao17], where the observer bandwidth
${\omega _o}$
is the only tunable parameter.
The linear feedback control law is adopted as follows:

Thus, the closed-loop equation can be expressed as follows:

with
${X_c} = {\left[ {\alpha, q,{z_1},{z_2},{z_3}} \right]^{\rm{T}}}$
,


Next, we designed a suitable controller parameter gain-scheduling algorithm to tune
${k_p}$
and
${k_d}$
for different operating conditions.
4.2 Guardian map theory and stability analysis
Guardian maps are scalar maps that project a set of real matrices of order
$N$
onto a region of a complex plane. Let
$\nu $
map
${{\rm{R}}^{n \times n}}$
onto the entire complex plane
$\mathbb{C}$
, and
${\rm{\Omega }}$
is a known region in the complex plane. The stability sets of interest are as follows [Reference Saydy, Tits and Abed27]:

Definition 3.
Let
$\nu $
map
${R^{n \times n}}$
onto
$\mathbb{C}$
. The map
$\nu $
guards
$S\left( \Omega \right)$
if the following equivalence holds for all
$A \in \mathop S\limits^\_ \left( \Omega \right)$
:

where
$\mathop S\limits^\_ \left( {\rm{\Omega }} \right)$
denotes the closure of the set
$S\left( {\rm{\Omega }} \right)$
, and
$\partial S\left( {\rm{\Omega }} \right)$
is the boundary of
$S\left( {\rm{\Omega }} \right)$
.
Figure 9 shows the GMs of some regions.
-
(1) Hurwitz stability (
${\rm{\Omega }} = {\mathbb{C}_ - }$ ): A GM is expressed as follows:

where
$ \otimes $
denotes the bialternate product [Reference Saydy, Tits and Abed27] (see Appendix).
-
(2) Stability margin: The open
$\sigma $ -shifted left half-plane region (
${\rm{R}}{{\rm{e}}_{}}\left( z \right) \lt \sigma $ ) is guarded by the following equation:

-
(3) The region enclosed by the circle centred at the origin of the complex plane with radius
${\omega _n}$ . A GM that corresponds to this conic damping ratio region is expressed as follows:

-
(4) A conic surface with an interior angle of 2
$\gamma $ , and a region with a damping ratio greater than
$\xi = {\rm{cos}}\gamma $ . The corresponding GMs are expressed as follows:


Figure 9. Guardian maps (GMs) for some stable regions.
The GMs for the stabilisation region of the flight control system considering the stable margin, natural frequency and damp ratio are as follows:

Based on these, a GM related to a specific region can be obtained. To analyse the largest variation range of the parameter that guarantees the stability of the parameterised matrices, we introduce Lemma 1 [Reference Saydy, Tits and Abed27].
Lemma 1.
When
$r$
is scalar and there is
${r_{min}} \le r \le {r_{max}}$
, for some value of
${r_0}$
such that
$A\left( {{r_0}} \right)$
is stable with respect to
$\Omega $
,
$\nu $
guards
$S\left( \Omega \right)$
. Therefore, the largest stable interval that stabilises
$A\left( r \right)$
with respect to
$\Omega $
and contains
${r_0}$
is
$r \in \left[ {{r^ - },{r^ + }} \right]$
, where

${r^ - }$
is the largest real root of
${\nu _{\rm{\Omega }}}\left[ {A\left( r \right)} \right] = 0$
on the interval
$\left[ {{r_{{\rm{min}}}},{r_0}} \right)$
(if it does not exist, then
${r^ - } = {r_{{\rm{min}}}}$
), and
${r^ + }$
is the smallest real root of
${\nu _{\rm{\Omega }}}\left[ {A\left( r \right)} \right] = 0$
on the interval
$\left[ {{r_0},{r_{{\rm{max}}}}} \right)$
(if it does not exist, then
${r^ + } = {r_{{\rm{max}}}}$
).
When
$r$
is scalar, given a stable region
${\rm{\Omega }}$
, the range of values of
$A\left( r \right)$
with respect to the region stable
$r$
can be obtained from Lemma 1 using GMs
$\nu $
.
We consider the stability of two-parameter polynomial families of real matrices (or polynomials) relative to a domain
${\rm{\Omega }}$
for which
$S\left( {\rm{\Omega }} \right)$
is endowed with a polynomic GM
${\nu _{\rm{\Omega }}}$

where
${r_1}$
and
${r_2}$
are the real parameters. To analyse the necessary and sufficient conditions for the stability of
$A\left( {{r_1},{r_2}} \right)$
over a specific rectangle-shaped domain
$\left( {{r_1},{r_2}} \right) \in \left[ {{\alpha _1},{\beta _1}} \right] \times \left[ {{\alpha _2},{\beta _2}} \right]$
Lemma 2 [Reference Saussié, Saydy, Akhrif and Berard39] is introduced. The aim is to find the largest stable open rectangle with a fixed side. First, let
${r^0} = \left( {r_1^0,r_2^0} \right)$
be a nominal parameter vector such that
$A\left( {r_1^0,r_2^0} \right)$
is
${\rm{\Omega }}$
stable. By freezing one of the parameters (
${r_2} = r_2^0$
), the largest (open) interval containing
$r_1^0$
,
${I_{{\rm{max}}}}\left( {r_2^0} \right) = \left( {r_1^ -, r_1^ + } \right)$
can be computed such that
$\forall {r_1} \in {I_{{\rm{max}}}}\left( {{r_1},r_2^0} \right) \in S\left( {\rm{\Omega }} \right)$
(Lemma 1). Let
${\nu _1}\left( {{r_2}} \right) = {p_{\left[ {{r_2}} \right]}}\left( {r_{1s}^ - } \right){p_{\left[ {{r_2}} \right]}}\left( {r_{1s}^ + } \right)$
, wherein
${p_{\left[ {{r_2}} \right]}}\left( {{r_1}} \right) = {\nu _{\rm{\Omega }}}\left[ {A\left( {{r_1},{r_2}} \right)} \right]$
represents a polynomial with
${r_1}$
as the variable for each
${r_2}$
.
Lemma 2.
Let
$A\left( {{r_1},{r_2}} \right)$
be expressed as in Equation (27), where
${r_{1,min}} \le {r_1} \le {r_{1,max}}$
,
${r_{2,min}} \le {r_2} \le {r_{2,max}}$
. Let
$A\left( {r_1^0,r_2^0} \right) \in S\left( \Omega \right)$
and let
${\nu _\Omega }$
be a GM for
$S\left( \Omega \right)$
. We choose
$\left[ {r_{1s}^ -, r_{1s}^ + } \right] \subset {I_{max}}\left( {r_2^0} \right)$
. We define
${\underline r _2} = sup\left\{ {{r_2} \in \left[ {{r_{2,min}},r_2^0} \right)\,:\,{\nu _1}\left( {{r_2}} \right) = 0} \right\}\left( {or - \infty ifnoneexist} \right)$
and
${\mathop r\limits^\_ _2} = inf\left\{ {{r_2} \in \left( {r_2^0,{r_{2,max}}} \right]\,:\,{\nu _1}\left( {{r_2}} \right) = 0} \right\}\left( {or + \infty ifnoneexist} \right)$
. The largest stability domain containing
$\left( {r_1^ -, r_1^ + } \right) \times r_2^0$
is then defined as
$\left[ {r_{1s}^ -, _{1s}^ + } \right] \times \Big( {\mathop r\limits_{ - 2}, {{\mathop r\limits^\_ }_2}} \Big)$
.
4.3 Controller parameter gain-scheduling algorithm based on GM
Parameter tuning based on GM is an offline tuning method that adjusts the control parameters by determining whether the closed-loop poles meet the performance requirements within the target region. The target region adopted herein is defined as Equation (25) in section 4.2, encompassing the constraints on the closed-loop poles residing in the left half of the complex plane.
Gain scheduling under all operating conditions is a 2D scheduling problem that considers both altitude and Mach number. The basic concept is as follows: The traversal is performed over the envelope of the Mach number
$\left[ {5,10} \right] \times \left[ {24,34} \right]$
km. A controller is first designed at the left boundary point of the domain (
$M = 5$
,
$h = 24{\rm{km}}$
, i.e.
${\delta _M} = - 1$
,
${\delta _h} = - 1$
). Then,
${\delta _h}$
is fixed as -1, and Lemma 1 is employed to calculate the variable range for which the controller can stabilise the system. Next, new controller parameters are searched at the right boundary point of the stable range, and the corresponding stable variable range is calculated. This process is iterated until the stable variable range covers the entire Mach interval. Finally, Lemma 2 is employed to determine the maximum stable altitude region. The process is iterated at the upper bound of the altitude interval to generate a set of controller gains that ensure stability across the full operational envelope.
The procedure is as follows and is illustrated in detail in Algorithm 1.
Algorithm 1 Mach-altitude adaptive gain-scheduling algorithm based on GM
Step 1: Initialisation
-
(1) Define the target stabilisation region
${{\rm{\Omega }}_t}$ and the corresponding GM
${\nu _{{{\rm{\Omega }}_t}}}$ .
-
(2) Specify the ranges of Mach numbers
$\left( {{\delta _{M,{\rm{min}}}},{\delta _{M,{\rm{max}}}}} \right)$ , and altitudes
$\left( {{\delta _{h,{\rm{min}}}},{\delta _{h,{\rm{max}}}}} \right)$ .
-
(3) Let
$i = 1$ ,
$\delta _h^i = {\delta _{h,{\rm{min}}}}$ .
Step 2: Controller parameter initialisation
Fix
$\delta _h^i$
. Let
$j = 1$
and
${\delta _{M0}} = {\delta _{M,{\rm{min}}}}$
. Find an initial controller parameter
${K_j} = \left[ {{k_p},{k_d}} \right]$
that ensures that the characteristic roots of
$A\left( {{\delta _{M0}},{\delta _{h0}},{K_j}} \right)$
lie within
${{\rm{\Omega }}_t}$
.
Step 3: Robustness analysis
-
(1) Calculate
${\nu _{{{\rm{\Omega }}_t}}}\left[ {A\left( {{\delta _M},\delta _h^i,{K_j}} \right)} \right] = 0$ and use Lemma 1 to calculate the maximum interval
$\left( {{\delta _{M,}}_j^ -, {\delta _{M,}}_j^ + } \right)$ for which
${\nu _{{{\rm{\Omega }}_t}}}\left[ {A\left( {{\delta _M},\delta _h^i,{K_j}} \right)} \right]$ is stable with respect to
${{\rm{\Omega }}_t}$ .
-
(2) If
$\delta _{M,j}^ + \geqslant {\delta _{M,{\rm{max}}}}$ , go to Step 6; otherwise, go to the next step with
${\delta _{M,j}} = \delta _{M,j}^ + $ .
Step 4: Find the central controller parameter
${K_{j + 1}}$
The central value of the stable region is obtained iteratively using a 2D search algorithm to determine the central controller parameters
${K_{j + 1}}$
, which provide the maximum margin. The maximum search times and the termination accuracy are specified as
${N_s}$
and
$\varepsilon $
, respectively.
-
(1) Let
${\delta _{M,j}} = \delta _{M,j}^ + $ ,
$m = 1$ and
${K^m} = {K_j}$ .
-
(2) While
$m \lt {N_s}$ , fix
${k_d}$ and use Lemma 1 to find the largest stability interval for
${k_p}$ as
$k_{}^{m + 1} = \left( {k_p^{m - } + k_p^{m + }} \right)/2$ .
-
(3) Fix
${k_p}$ and seek the largest stability interval for
${k_d}$ as
$k_d^{m + 1} = \left( {k_d^{m - } + k_d^{m + }} \right)/2$ .
-
(4) When
${K^{m + 1}} - {K^m} \le \varepsilon \left( {1 + {K^m}} \right)$ , go to Step 5; otherwise, let
$m = m + 1$ and go to Step 2.
-
(5)
${K_{j + 1}} = {K^{m + 1}}$ . Recalculate the GM
${\nu _{{{\rm{\Omega }}_t}}}\left[ {A\left( {{\delta _M},{\delta _h},{K_{j + 1}}} \right)} \right] = 0$ and seek the next stability interval
$\left( {{\delta _{M,}}_{j + 1}^ -, {\delta _{M,}}_{j + 1}^ + } \right)$ containing
${\delta _{M,j}}$ .
-
(6) If
$\delta _{M,j + 1}^ + \geqslant {\delta _{M,{\rm{max}}}}$ , proceed to Step 5; otherwise, repeat Step 4 with
$j = j + 1$ .
Step 5: Determine the range of
${\delta _h}$
$\left( {\mathop \delta \limits_{ - h}^i, \mathop \delta \limits_h^{ - i} } \right)$
when
${K_j}$
stablises
$A\left( {{\delta _M},\delta _h^i,{K_j}} \right)$
with respect to
${{\bf{\Omega }}_t}$
-
(1) Use Lemma 2 to find the largest stability domain
$\left( {{\delta _{M,}}_j^ -, {\delta _{M,}}_j^ + } \right) \times \left( {\underline \delta _{h,j}^i,\bar \delta _{h,j}^i} \right)$ containing.
-
(2) Take their largest common intersection as
$\left( {\mathop \delta \limits_{ - h}^i, \mathop \delta \limits_h^{ - i} } \right)$ .
Step 6: Terminating indicator
If
$\bar \delta _h^i \gt {\delta _{h,{\rm{max}}}}$
, the end. Otherwise let
$\delta _h^{i + 1} = \bar \delta _h^i$
,
$i = i + 1$
, and return to Step 2.
This process generates a set of controller parameters
$\left\{ {{K_1},{K_2}, \ldots {K_j}} \right\}$
and the corresponding range sets
$\left\{ {\left[ {{\delta _{M,{\rm{min}}}},\delta _{M,1}^ + } \right),...,\left( {\delta _{M,j}^ -, \delta _{M,j}^ + } \right),...,\left( {\delta _{M,j}^ -, {\delta _{ma,{\rm{max}}}}} \right]} \right\} \times \left( {\underline \delta _h^i,\bar \delta _h^i} \right)$
that meet the stability requirements. An analytical expression for the controller parameters
$K$
as a function of the variable (
${\delta _M}$
,
${\delta _h}$
) can be derived by curve fitting.
Remark: Assuming the length of each side of a rectangle is constrained to be less than the Lipschitz constant
$L$
, it is guaranteed that smaller rectangles can adequately cover the entire envelope.
5.0 Results and discussions
The simulation parameters are listed in Table 2. The controller design criteria consider the stability margin, the short-period damping ratio and the natural frequency. To ensure good stationarity, rapidity and high precision, the constraints of assigning the closed-loop pole in the target stability region are a stability margin
$\left| \sigma \right| \geqslant 4$
and a damping ratio
$\xi \geqslant 0.707$
, and a natural frequency
${\omega _n} \le 16$
.
Table 2. The parameters in the simulation

Figure 10 shows the variation in the control efficiency within the flight envelope. The control efficiency increased from approximately 2 to 12, highlighting the necessity of gain-scheduled control. As shown in Fig. 11 when
${\delta _h} = 1$
, there is almost no intersection in the case of
${\delta _M} = - 1$
and
${\delta _M} = 1$
, indicating that a single controller cannot meet all the conditions. These results demonstrate that the gain-scheduling controller is vital.

Figure 10. Variation of control efficiency within the flight envelope.

Figure 11. Stabilising gain regions for different values of
${\delta _M} = \left\{ {1, - 1} \right\}$
when
${\delta _h} = 1$
.
The algorithm described in section 4.3 was employed to calculate gain scheduling with Mach and altitude. First, the initial controller
${K_1}$
at the starting point (
$M = 5$
,
$h = 24$
km, i.e.
${\delta _M} = - 1$
,
${\delta _h} = - 1$
) was designed to ensure that the closed-loop poles were within the target region. According to Lemma 1,
${K_1}$
can cover a Mach number range of (−1.489, 0.022). With
${\delta _M}$
fixed at the right boundary point 0.022, the 2D search algorithm obtains the centre value of the stable region as the new controller
${K_2}$
(Fig. 12). Algorithm 1 was applied and iterated to obtain the values of 49 groups of controllers (see Table A1 in the Appendix for details). The 49 groups of controllers can cover all the conditions.
${k_p}$
and
${k_d}$
can be obtained by fitting
${\delta _M}$
and
${\delta _h}$
as variables in Equation (28), as shown in Fig. 13.


Figure 12. Procedure for seeking primary and central controllers in a 2D subspace.

Figure 13. Variation of the controller parameters
${k_p}$
and
${k_d}$
with
${\delta _M}$
and
${\delta _h}$
.
For the intervals of
${\delta _M} \in \left[ { - 1,1} \right],{\delta _h} \in \left[ { - 1,1} \right]$
, values were taken every 0.2, resulting in a total of 121 state points. Under different Mach and flight altitudes, the control law Equation (28) was applied to the closed-loop state Equation (18). Figure 14 shows the pole distributions of the short-period modes corresponding to the short-period modes under the given conditions. The open-loop poles have a low damping ratio and are close to the imaginary axis, whereas the closed-loop poles are in the target region, which satisfies the flight control performance index.

Figure 14. Open and closed-loop pole distributions for the full flight envelope.
5.1 Simulation of step response and climb phase
To verify the effectiveness of the proposed method, we compared it with the traditional ADRC method and GM-LQR developed by Xiao et al. [Reference Liu, Chen, Chen and Liu30]. The proposed gain-scheduling algorithm based on GM is called ‘GM-ADRC’, and that applied to the PID control is termed ‘GM-PID’. The traditional ADRC algorithm (hereafter ADRC) does not perform gain scheduling of
${k_p}$
and
${k_d}$
but only adjusts the corresponding
${b_0}$
according to different working conditions. The gain value of ADRC was taken from a point [
${k_p}$
,
${k_d}$
] = [102.38, 17.84] of the fitted surface.
Test 1: Step response test under all conditions
Given a step signal with an amplitude of 3° as the angle-of-attack tracking command signal, the angle-of-attack response curve is shown in Fig. 15. The integral of the time absolute error (ITAE), integral of the absolute error (IAE), and settling time (
${T_s}$
) were employed as performance indicators to compare the results of the algorithms (Table 3).
Table 3. Performance index comparison of Experiment 1


Figure 15.
$\alpha $
response under different operating conditions.
Both GM-ADRC and ADRC could quickly follow the given value, demonstrating good response characteristics and indicating the feasibility of ADRC control. The GM-ADRC was faster than the standard ADRC and exhibited superior performance indicators. GM-ADRC exhibited a lower ITAE than ADRC, indicating that it not only responded faster but also had a lower overall error over time. The lower IAE value for GM-ADRC also indicates that the magnitude of the cumulative error throughout the response is reduced. Finally, the lower
${T_s}$
value for GM-ADRC confirms that the algorithm reaches the desired output level sooner, indicating that the proposed algorithm exhibited faster adjustment to changes and disturbances.
Figures 16 and 17 show the response curves of the pitch angle rate
$q$
and the elevator deflection angle
${\delta _e}$
at different Mach numbers, respectively.
${\delta _e}$
and
$q$
are stabilised, and
$q$
is also stabilised at approximately
$0\left( {{\rm{deg}}/{\rm{s}}} \right)$
, indicating that the vehicle finally maintains a stable attitude. Therefore, the proposed controller exhibits a good control effect.

Figure 16.
$q$
response under different operating conditions.

Figure 17.
${\delta _e}$
response under different operating conditions.
Test 2: Gain-scheduling control in the power climb stage
To verify the effectiveness of the algorithm in trajectory flight, we designed guidance instructions and employed the proposed gain-scheduling control in the powered climb stage. In this phase, the climbing plus and constant altitude cruise flight state should be achieved. The guidance instruction for the angle-of-attack was obtained by PID control of the actual altitude error, which is used as the tracking instruction signal for the actual attitude control. The altitude guidance instructions are as follows:

where
$a = \frac{{{h_0} - {h_c}}}{{\left( {{t_f} - {t_m}} \right){t_m} + {{\left( {{t_f} - {t_m}} \right)}^2}/2}},b = a{\left( {{t_f} - {t_m}} \right)^2}/4 + {h_0},{t_m} = {t_2} - {t_1}$
.
The angle-of-attack guidance instructions are as follows:

with
${\rm{sat}}\left( {{\Delta}h,500} \right) = \begin{cases} {{\Delta}h,\left| {{\Delta}h} \right| \le 500} \\[3pt] {500,{\Delta}h \gt 500} \\[3pt] { - 500,{\Delta}h \lt - 500} \end{cases}\!\!\!\!\!\!, $
${\Delta}h = {h_r} - h$
is the height error,
${t_{h = 500}}$
is the first time
${\Delta}h$
=500m, and
${\Delta}\dot h$
is the derivative of the height error.
The above angle-of-attack of the guidance command was adopted, and the parameters were set as follows:
${t_1} = 150$
,
${t_2} = 250$
,
${t_f} = 400$
,
${h_0} = 22{\rm{km}}$
,
${h_c} = 33.5{\rm{km}}$
,
${k_{ph}} = 0.0074$
,
${k_{dh}} = 0.0573$
, and
${k_{ih}} = 1.146 \times {10^{ - 4}}$
. Figure 18 shows the obtained altitude and velocity of the particle trajectory along with the corresponding state variables. The results show that the proposed gain-scheduling strategy can be employed in the climbing stage of aircraft. Furthermore, we compared the angle-of-attack tracking effects and performance indicators of GM-ADRC, ADRC, GM-PID and GM-LQR [Reference Liu, Chen, Chen and Liu30] (Fig. 19 and Table 4). All algorithms could follow the angle-of-attack instructions, but GM-ADRC exhibited the lowest ITAE (2.0329) and IAE (0.0621), indicating better tracking accuracy.

Figure 18. Variation of altitude,
${\delta _e}$
, and
$\gamma $
with time.

Figure 19. Comparison of the angle-of-attack
$\alpha $
tracking for the different algorigthms.
Table 4. Performance index comparison of Experiment 2

5.2 Robustness analysis
To account for uncertainties in the vehicle model, random variations of up to 20% were introduced to each aerodynamic parameter, and Monte Carlo simulations were performed 1,000 times to evaluate the robustness of the control system. Monte Carlo simulations were performed 1,000 times to evaluate the robustness of the control system. Figure 20 shows the angle-of-attack
$\alpha $
and elevator deflection
${\delta _e}$
curves of the proposed algorithm. The results demonstrate that within the specified uncertainty range, the angle-of-attack consistently follows the commanded signal, and the controller exhibits robust and reliable performance in maintaining control effectiveness.

Figure 20. Change curve of
$\alpha $
and
${\delta _e}$
with 20% aerodynamic parameters bias.
To evaluate the aircraft’s response to wind gusts, we simulated the climb phase under gust disturbance. The Horizontal Wind Model (HWM) [Reference Drob, Emmert, Meriwether, Makela, Doornbos, Conde and Klenzing40,Reference Drob, Emmert, Crowley, Picone, Shepherd, Skinner and Vincent41], which is widely recognised and adopted by the U.S. Department of Defense [Reference Madden42], was employed to mathematically represent continuous gusts. This model characterises the horizontal wind field in the stratosphere based on changes in altitude and latitude. The tangential wind velocity is expressed as follows:

where
${v_{{\rm{wind}}}}$
is the tangential wind speed at altitude
$h$
and latitude
$\varphi $
,
${h_0}$
is the reference altitude (set to 24 km herein), and
$\varphi $
was set to 45°. The baseline wind speed,
${v_0}\left( \varphi \right)$
, is defined as
$10 + 20 \cdot \left| {{\rm{sin}}\left( {\varphi \cdot \pi /180} \right)} \right|$
. The coefficients
$a\left( \varphi \right)$
and
$b\left( \varphi \right)$
, which represent the linear and quadratic altitude variations, are defined as
$a\left( \varphi \right) = 2 + 0.1 \cdot \left| {{\rm{sin}}\left( {\varphi \cdot \pi /180} \right)} \right|$
and
$b\left( \varphi \right) = 0.05 \cdot {\rm{cos}}\left( {\varphi \cdot \pi /180} \right)$
, respectively.
Figure 21 shows the tangential wind speed profile. The wind disturbance was applied between 100 and 200s to evaluate the system’s robustness under realistic atmospheric conditions. Figure 22 shows the angle-of-attack
$\alpha $
response. The wind caused an initial sharp deviation in the angle-of-attack
$\alpha $
response, but it gradually stabilised with a minor error. As shown in Table 5, the ITAE and IAE values demonstrate the robustness and accuracy of the proposed controller compared with the GM-PID controller, particularly under wind disturbance. Figure 23 compares Mach-altitude flight envelopes with and without wind disturbance. Overall, the system maintained stability and consistency under disturbances, demonstrating its robustness and reliability in overcoming wind disturbances.

Figure 21. Wind gust velocity profile.

Figure 22. Angle-of-attack (
$\alpha $
) response under wind gust disturbance.
Table 5. Performance index comparison under wind disturbance


Figure 23. Effects of wind disturbance on flight envelopes.
6.0 Conclusions
In this study, we developed an automatic gain-scheduling algorithm based on the GM theory using a more effective LPV model of aircraft as the object. The application of LADRC to a hypersonic vehicle leverages the advantages of traditional PID control, and it exhibits good control performance under severe model uncertainties. The proposed controller parameter gain-scheduling algorithm based on GMs eliminates the need to design a controller at multiple state points. This approach reduces the workload and streamlines the design process. The proposed algorithm meets the performance requirements of the entire envelope and outperforms the traditional ADRC, GM-PID and GM-LQR algorithms. To validate the robustness of the proposed algorithm, we performed Monte Carlo simulations of the aerodynamic parameters under wind disturbance.
Acknowledgements
This work was funded by the National Nature Science Foundation of China, Grant Nos. 62473209, 62073177 and 61973175. This manuscript builds upon the authors’ prior conference paper, ‘Active Disturbance Rejection Gain-Scheduling Control of Hypersonic Vehicle Based on Guardian Maps,’ presented at the 2024 IEEE 13th Data-Driven Control and Learning Systems Conference (DDCLS), with several remarkable improvements and extensions.
Competing interests
The authors declare no competing interests relevant to the content of this article.
Ethical approval
This study is related to the improvement and application innovation of gain-scheduling algorithms; thus, it does not involve ethical issues. All authors are aware of this paper and agree to its submission.
Research transparency
No datasets were analysed during this study. The data generated in this study are available from the corresponding author upon reasonable request.
Appendix
The coefficient expression used in this paper is as follows:




Bilaternate Product
Let A and B be
$n \times n$
matrices. Let
${V^n}$
be the
$\frac{1}{2}n\left( {n - 1} \right)$
-tuple consisting of pairs of integers
$\left( {p,q} \right)$
,
$p = 2,3 \ldots n$
,
$q = 1, \ldots, p - 1$
, listed lexicographically. That is,
${V^n} = \left[ {\left( {2,1} \right),\left( {3,1} \right),\left( {3,2} \right),\left( {4,1} \right),\left( {4,2} \right),\left( {4,3} \right), \ldots \left( {n,n - 1} \right)} \right].$
Denote by
$V_i^n$
the
$i$
th entry of
${V^n}$
. Denote

where the dependence of
$f$
on
$A$
and
$B$
is kept implicit for simplicity. The bialternate product of
$A$
and
$B$
, denoted by
$A \otimes B$
, is a
$\frac{1}{2}n\left( {n - 1} \right)$
-dimensional square matrix, for which the
$ij$
th entry is given by
${\left( {A \otimes B} \right)_{ij}} = f\left( {V_i^n,V_j^n} \right)$
.
Table A1. Controller parameters and stabilisation intervals for Mach number changes
