šÆ Problem Statement
Consider a quadrotor UAV operating under intermittent feedback conditions where state measurements are periodically unavailable.
The challenge is to design an adaptive controller that maintains stability and tracking performance during both feedback-available
and feedback-lost periods while learning unknown system dynamics.
Key Innovation: Lyapunov-based Deep Neural Network (LbDNN) adaptation with mathematically rigorous stability guarantees
under intermittent feedback loss scenarios.
š Nonlinear System Dynamics
Second-Order Nonlinear System:
$$\ddot{x} = f(x,\dot{x}) + g(x,\dot{x})u$$
State Definitions:
⢠$x, \dot{x} \in \mathbb{R}^{n}$: States with available measurements
⢠$\ddot{x} \in \mathbb{R}^{n}$: Unknown state-derivative
⢠$f: \mathbb{R}^{n} \times \mathbb{R}^{n} \rightarrow \mathbb{R}^{n}$: Unknown continuously differentiable drift function
⢠$g: \mathbb{R}^{n} \times \mathbb{R}^{n} \rightarrow \mathbb{R}^{n \times m}$: Known locally Lipschitz control effectiveness matrix
⢠$u \in \mathbb{R}^{m}$: Control input
For Quadrotor Application: The state vector $x = [x, y, z, \phi, \theta, \psi]^T \in \mathbb{R}^6$
represents position and orientation, with control inputs $u = [F_x, F_y, F_z, \tau_\phi, \tau_\theta, \tau_\psi]^T \in \mathbb{R}^6$
representing forces and torques applied to achieve desired trajectory tracking.
ļæ½ Tracking Error Formulation
Tracking Error Definition:
$$e \triangleq x - x_d(t)$$
Auxiliary Error (Sliding Variable):
$$r \triangleq \dot{e} + \alpha_1 e$$
Error Dynamics:
$$\dot{r} = f(x,\dot{x}) + g(x,\dot{x})u - \ddot{x}_d(t) + \alpha_1(r - \alpha_1 e)$$
Design Parameter: $\alpha_1 \in \mathbb{R}_{>0}$ is a constant control gain that determines the convergence rate
of the tracking error dynamics and affects the sliding surface slope.
āļø Control Law Design
Control Input Design:
$$u = g^+(x,\dot{x})\left(\ddot{x}_d(t) - (\alpha_1 + k_r)r + (\alpha_1^2 - 1)e - \Phi(X,\hat{\theta})\right)$$
Control Parameters:
⢠$g^+(x,\dot{x})$: Moore-Penrose pseudoinverse of control effectiveness matrix
⢠$k_r \in \mathbb{R}_{>0}$: Sliding mode gain
⢠$\Phi(X,\hat{\theta})$: DNN approximation of unknown dynamics
⢠$X = [x^T, \dot{x}^T]^T \in \mathbb{R}^{2n}$: Augmented state vector
Control Strategy: The control law combines feedforward compensation (desired acceleration),
feedback stabilization (sliding mode terms), and adaptive uncertainty compensation (neural network approximation).
š Dynamic State-Derivative Observer
Observer Dynamics:
$$\dot{\hat{r}} = g(x,\dot{x})u - \ddot{x}_d(t) + \alpha_1(r - \alpha_1 e) + \hat{f} + \alpha_2\tilde{r}$$
$$\dot{\hat{f}} = k_f(\dot{\tilde{r}} + \alpha_2\tilde{r}) + \tilde{r}$$
Observer Error Dynamics:
$$\dot{\tilde{r}} = \tilde{f} - \alpha_2\tilde{r}, \qquad \dot{\tilde{f}} = \dot{f} - k_f\tilde{f} - \tilde{r}$$
Observer Variables:
⢠$\hat{r}, \hat{f} \in \mathbb{R}^n$: Observer estimates of $r$ and $f$
⢠$\tilde{r} \triangleq r - \hat{r}$, $\tilde{f} \triangleq f(x,\dot{x}) - \hat{f}$: Observer errors
⢠$\alpha_2, k_f \in \mathbb{R}_{>0}$: Constant observer gains
⢠$\dot{f} \triangleq \frac{\partial f}{\partial x}\dot{x} + \frac{\partial f}{\partial \dot{x}}\ddot{x}$: Time derivative of drift function
Implementation Note: Since $\dot{\tilde{r}}$ is unknown, equation (2) is implemented by integration:
$$\hat{f}(t) = \hat{f}(t_0) + k_f\tilde{r}(t) - k_f\tilde{r}(t_0) + \int_{t_0}^t (k_f\alpha_2 + 1)\tilde{r}(\tau)d\tau$$
š§ Composite Adaptation Law
Prediction Error Design:
$$E \triangleq \hat{f} - \Phi(X,\hat{\theta})$$
Composite Least Squares Adaptation:
$$\dot{\hat{\theta}} = \text{proj}\left(-k_{\hat{\theta}}\Gamma(t)\hat{\theta} + \Gamma(t)\Phi'^T(X,\hat{\theta})(r + \alpha_3 E)\right)$$
Time-Varying Adaptation Gain:
$$\frac{d}{dt}\Gamma^{-1} = \begin{cases}
-\beta(t)\Gamma^{-1} + \Phi'^T(X,\hat{\theta})\Phi'(X,\hat{\theta}), & \text{if } \lambda_{\Gamma,\min} < \lambda_{\min}(\Gamma) \text{ and } \lambda_{\max}(\Gamma) < \lambda_{\Gamma,\max} \\
0_{p \times p}, & \text{otherwise}
\end{cases}$$
Key Innovation: The prediction error $E$ uses the dynamic state-derivative estimator $\hat{f}$
instead of traditional approaches, enabling composite adaptation for non-linearly parameterized (NIP) neural networks.
Adaptation Parameters:
⢠$\alpha_3, k_{\hat{\theta}} \in \mathbb{R}_{>0}$: Constant adaptation gains
⢠$\Phi'(X,\hat{\theta}) \triangleq \frac{\partial \Phi(X,\hat{\theta})}{\partial \hat{\theta}} \in \mathbb{R}^{n \times p}$: Jacobian matrix
⢠$\text{proj}(\cdot)$: Continuous projection operator ensuring $\hat{\theta}(t) \in \mathcal{B}_{\bar{\theta}}$
⢠$\beta(t) \triangleq \beta_0\left(1 - \frac{\|\Gamma(t)\|}{\varkappa_0}\right)$: Bounded forgetting factor
ļæ½ļø Intermittent Feedback Framework
Time Partition:
⢠Feedback Available: $t \in [t_{2i}, t_{2i+1})$, $i \in \mathbb{N}$
⢠Feedback Lost: $t \in [t_{2i+1}, t_{2i+2})$, $i \in \mathbb{N}$
⢠Total feedback loss duration: $\sum_i (t_{2i+2} - t_{2i+1}) \leq T_{\max}$
Control During Feedback Loss:
⢠Neural network weights frozen: $\dot{\hat{\theta}} = 0$
⢠Observer adaptation suspended: $\dot{\hat{r}} = 0$, $\dot{\hat{f}} = 0$
⢠Open-loop control using last known states and frozen DNN
During feedback-available periods, the controller has access to state measurements $x(t)$, $\dot{x}(t)$ and performs
active adaptation. During feedback-lost periods, the controller operates open-loop using the last known state
information and frozen learned dynamics.
šŖļø Turbulent Nonlinearity Modes
The simulation incorporates nine distinct turbulent nonlinearity modes that represent real-world aerodynamic and
mechanical phenomena that significantly challenge quadrotor control systems. Each mode introduces specific
state-dependent disturbances that test the adaptive controller's robustness.
š¬ļø 1. Aerodynamic Stall & Rotor Interference
Mathematical Model:
$\text{AoA}_{\text{eff}} = \arctan\left(\frac{\dot{z}}{\sqrt{\dot{x}^2 + \dot{y}^2 + 0.001}}\right)$
$S_{\text{stall}} = 1 - 0.8 \exp\left(-\frac{(\text{AoA}_{\text{eff}} - 0.3)^2}{0.15^2}\right)$
$F_{\text{interference}} = 0.1 \sin(2\phi) \cos(3\theta) |\mathbf{v}|$
Physical Origin: When the effective angle of attack exceeds critical values (~17°), airflow separates
from rotor blades causing dramatic lift loss. Rotor wake interference creates asymmetric forces that depend on
attitude angles and velocity magnitude. This nonlinearity is particularly dangerous during aggressive maneuvers.
š 2. Ground Effect
Mathematical Model:
$\text{GE}_{\text{factor}} = \left(\frac{D_{\text{rotor}}}{h_{\text{ground}}}\right)^2 \exp\left(-\frac{h_{\text{ground}}}{D_{\text{rotor}}}\right)$
$F_{\text{lift,aug}} = \text{GE}_{\text{factor}} \cdot 0.15 \tanh(5\dot{z})$
$\tau_{\text{coupling}} = 0.05 \cdot \text{GE}_{\text{factor}} [\sin(\phi), \sin(\theta), 0]^T$
Physical Origin: When operating near the ground (typically $h < 2D_{\text{rotor}}$), the
downwash reflects off the surface creating a "cushioning" effect that increases effective lift. This exponentially
decaying phenomenon also induces attitude-dependent moments due to asymmetric pressure distribution.
š 3. Vortex Ring State
Mathematical Model:
$\text{VRS}_{\text{condition}} = (\dot{z} < -2.0) \land (|\mathbf{v}_{\text{horizontal}}| < 1.5)$
$I_{\text{VRS}} = \exp\left(-\frac{(|\mathbf{v}_{\text{horizontal}}| - 0.5)^2}{0.3^2}\right)$
$F_{\text{VRS}} = -I_{\text{VRS}} \cdot 8.0 \sin(0.5t) \hat{\mathbf{z}}$
Physical Origin: During rapid descent with low forward speed, the rotor enters its own wake,
creating a toroidal vortex that severely disrupts lift generation. This dangerous flight condition causes
oscillatory forces and loss of control authority, requiring immediate recovery maneuvers.
šļø 4. Structural Flexibility
Mathematical Model:
$F_{\text{flex},x} = 0.1 \sin(\phi) (\dot{x}^2 + \dot{y}^2)$
$F_{\text{flex},y} = 0.08 \cos(\theta) \dot{z} \dot{\phi}$
$\tau_{\text{flex},\psi} = 0.12 (\phi^2 + \theta^2) \text{sign}(\dot{\psi})$
Physical Origin: Real quadrotor frames exhibit elastic deformation under aerodynamic loads
and inertial forces. This creates complex coupling between translational and rotational dynamics, where
structural bending produces position-dependent attitude disturbances and vice versa.
š 5. Actuator Nonlinearity & Battery Degradation
Mathematical Model:
$\beta_{\text{battery}}(t) = 1 - \alpha_{\text{deg}} \tanh(t/30)$
$\xi_{\text{motor}} = 0.1(1 - \beta_{\text{battery}})$
$F_{\text{nonlin}} = \xi_{\text{motor}} [\sin(2t)\dot{x}, \cos(2.3t)\dot{y}, \sin(1.7t)\dot{z}]^T$
Physical Origin: Electric motors exhibit nonlinear torque-speed characteristics, especially
under varying voltage conditions. Battery degradation reduces available power, causing time-varying actuator
effectiveness and introducing harmonic disturbances in the control response.
š 6. Atmospheric Chaos & Multi-Scale Turbulence
Mathematical Model:
$W_{\text{large}} = I_{\text{turb}} \sin(0.05t) \cos(0.3x + 0.2y)$
$W_{\text{small}} = 0.5 I_{\text{turb}} \sin(2.1t + \sigma_{\text{state}}) \cos(5.7t)$
$W_{\text{gust}} = I_{\text{turb}} \exp\left(-\frac{(t \bmod 20 - 10)^2}{9}\right)$
Physical Origin: Atmospheric turbulence exhibits complex multi-scale behavior with large eddies
(building-scale), small eddies (vehicle-scale), and intermittent gust fronts. This creates spatially and
temporally correlated disturbances that challenge predictive control algorithms.
ā” 7. Reynolds Number-Dependent Drag
Mathematical Model:
$\text{Re} = \frac{|\mathbf{v}| L_{\text{char}}}{\nu_{\text{air}}}$
$C_D(\text{Re}) = 0.1 + \frac{0.4}{1 + \exp(-(Re - 1000)/200)}$
$\mathbf{F}_{\text{drag}} = -C_D \cdot |\mathbf{v}| \mathbf{v} \odot \text{sign}(\mathbf{v})$
Physical Origin: The aerodynamic drag coefficient varies significantly with Reynolds number,
transitioning from laminar to turbulent flow regimes. This velocity-dependent nonlinearity affects both
the magnitude and direction of aerodynamic forces acting on the vehicle.
š 8. Propwash-Induced Coupling
Mathematical Model:
$v_{\text{wash}} = 5.0 + 2.0|\dot{z}|$
$F_{\text{wash},x} = 0.05 v_{\text{wash}} \sin(2\phi) \cos(\psi)$
$F_{\text{wash},y} = 0.05 v_{\text{wash}} \sin(2\theta) \sin(\psi)$
$F_{\text{wash},z} = -0.1 v_{\text{wash}} (1 + 0.3\cos\phi \cos\theta)$
Physical Origin: The high-velocity downwash from rotors creates attitude-dependent forces that
couple translational and rotational dynamics. These propeller-induced effects vary with rotor disk loading
and attitude angles, creating cross-axis disturbances during maneuvering.
š§² 9. Magnetic Interference & Electronic Coupling
Mathematical Model:
$B_{\text{field}}(t) = 0.02(1 + 0.5\sin(0.1t))$
$|\boldsymbol{\omega}| = \sqrt{\dot{\phi}^2 + \dot{\theta}^2 + \dot{\psi}^2}$
$F_{\text{mag}} = B_{\text{field}} [\sin\phi (x^2+y^2)/10, \cos\theta \cdot z \cdot \text{sign}(\dot{\psi}), 0]^T$
$\tau_{\text{mag}} = B_{\text{field}} \tanh(|\boldsymbol{\omega}|) \cdot 0.1$
Physical Origin: Electric motors and power electronics generate time-varying magnetic fields
that interact with the Earth's magnetic field and metallic structures. These electromagnetic effects couple
with vehicle dynamics, especially during high angular rate maneuvers, causing subtle but persistent disturbances.
š” Collective Impact on Adaptive Control
These nine turbulent nonlinearity modes can be selectively activated to test different aspects of the
adaptive controller's robustness. When combined, they create a highly challenging flight environment
that requires sophisticated adaptation mechanisms.
- State-Dependent Coupling: Most effects depend on current position, velocity, and attitude states
- Multi-Time-Scale Dynamics: Effects range from fast oscillations to slow drift phenomena
- Cross-Axis Coupling: Disturbances in one axis affect control authority in others
- Unknown Parameters: Physical constants vary with environmental conditions and aging