Fact-checked by Grok 2 weeks ago

Feedback linearization

Feedback linearization is a method in nonlinear control theory that uses state feedback and diffeomorphic coordinate transformations to algebraically cancel the nonlinearities in a system's dynamics, thereby rendering an equivalent linear system to which standard linear control techniques can be applied. The technique primarily applies to affine nonlinear systems of the form \dot{x} = f(x) + g(x)u with output y = h(x), where exact linearization requires the system to satisfy specific geometric conditions, such as the bracket distribution being involutive and the relative equaling the system dimension for full state . Two main approaches exist: input-output linearization, which focuses on linearizing the map from input to output by repeated differentiation until the control input appears explicitly, often resulting in a chain of integrators for systems with relative r < n; and state-space (or full) linearization, which transforms the entire state dynamics into a controllable linear form \dot{z} = Az + Bv via feedback u = \alpha(x) + \beta(x)v and state transformation z = \Phi(x), provided the controllability matrix has full rank. These transformations rely on the existence of a nonsingular diffeomorphism and are local or global depending on the system's properties. Originally developed in the early 1980s through differential geometric methods, feedback linearization has roots in works by researchers like Jakubczyk and Respondek (1980) on feedback equivalence and Hunt, Su, and Meyer (1983) on design applications, with comprehensive treatments later provided by Isidori in his 1995 book Nonlinear Control Systems. Its advantages include global stability guarantees for certain systems (e.g., minimum-phase plants) and simplified controller design, such as pole placement on the linear equivalent, but it demands precise knowledge of the model and can be sensitive to uncertainties or unmodeled dynamics. Applications span mechanical systems like robot manipulators—where nonlinear terms from inertia and Coriolis forces are canceled to yield simple double-integrator dynamics—and process control in chemical engineering, aerospace, and power systems.

Overview

Definition and Motivation

Feedback linearization is a control technique for nonlinear systems that employs state feedback and a nonlinear change of coordinates to transform the system's dynamics into an equivalent linear and controllable form, thereby allowing the application of well-established linear control methods such as pole placement or linear optimal control. This transformation achieves exact cancellation of the nonlinearities, rather than relying on local approximations, and is particularly suited to systems where the full state is measurable. The primary motivation for feedback linearization arises from the limitations of traditional linearization techniques, such as , which provide accurate models only in the vicinity of specific operating points and often fail to capture the global behavior of highly nonlinear systems. In applications like , where manipulator dynamics involve complex couplings and varying payloads, or aerospace systems, such as with significant aerodynamic nonlinearities, linear approximations can lead to poor performance or instability over wide operating ranges. By exactly compensating for these nonlinearities, feedback linearization enables robust and precise control design that performs effectively across the system's operational envelope. Consider a nonlinear system in affine form, described by the state-space equations \dot{x} = f(x) + g(x)u and output y = h(x), where x \in \mathbb{R}^n is the state vector, u \in \mathbb{R} is the scalar input, f and g are smooth vector fields, and h is a smooth scalar function. A static state feedback law of the form u = \alpha(x) + \beta(x)v, with \alpha and \beta appropriately chosen nonlinear functions and v as the new linear input, is designed to yield a linearized input-output relationship, transforming the system into a chain of integrators or a controllable linear canonical form. One key advantage of feedback linearization is the potential for global linearization when the system satisfies necessary structural conditions, such as having a well-defined relative degree equal to the system order, which simplifies controller synthesis and allows for straightforward assignment of desired closed-loop poles. This approach not only enhances tracking accuracy and disturbance rejection but also facilitates the integration with other techniques, like observer design for unmeasured states, in practical implementations.

Historical Context

Feedback linearization emerged in the 1980s as a key technique within nonlinear control theory, building on foundational applications of differential geometry to dynamical systems developed in the preceding decade. This approach addressed the limitations of linear control methods for inherently nonlinear processes, such as those in aerospace and robotics, by seeking transformations that render nonlinear dynamics equivalent to linear ones via state feedback and coordinate changes. Early geometric perspectives were advanced by Roger W. Brockett, who in 1978 introduced concepts of feedback invariants that highlighted intrinsic nonlinear structures resistant to linearization, laying groundwork for analyzing system equivalence under feedback. Central to the development were contributions from , whose work in the 1980s on normal forms for nonlinear systems provided a structured framework for decomposing dynamics into linearizable and internal components, enabling precise feedback designs. A seminal advancement came in 1983 with the collaboration of and Isidori, who demonstrated conditions under which nonlinear systems could be linearized through output injection, establishing feedback equivalence to linear models and influencing observer design for state estimation. Key milestones included extensions to multivariable systems, with Henk Nijmeijer in 1985 developing methods for MIMO feedback linearization that generalized relative degree concepts and decoupling strategies to multi-input scenarios, broadening applicability to complex industrial processes. By the 1990s, the technique evolved from theoretical constructs to practical implementations, incorporating robustness enhancements and real-time computations in areas like robotic manipulators and chemical process control, as evidenced in early applications documented in control literature. The advent of feedback linearization marked a paradigm shift in nonlinear control, moving away from Lyapunov-based stability analysis toward differential geometric tools that emphasized structural properties and exact transformations. This transition facilitated integrations with adaptive and robust control paradigms, profoundly impacting fields such as adaptive control by providing a pathway to handle parametric uncertainties alongside nonlinearities.

Mathematical Prerequisites

Lie Derivatives

In nonlinear control theory, the Lie derivative serves as a key differential geometric tool for quantifying the rate of change of functions or vector fields along the direction specified by a given vector field, enabling the analysis of nonlinear dynamical systems. For a smooth scalar function h: \mathbb{R}^n \to \mathbb{R} and a smooth vector field f: \mathbb{R}^n \to \mathbb{R}^n, the Lie derivative of h along f, denoted L_f h, is defined as the directional derivative L_f h(x) = \frac{\partial h(x)}{\partial x} f(x) = \nabla h(x) \cdot f(x). This expression represents the instantaneous variation of h in the direction of f at point x. Higher-order Lie derivatives are obtained iteratively as L_f^k h(x) = L_f (L_f^{k-1} h(x)), where L_f^0 h(x) = h(x), allowing for successive differentiations along the vector field. The Lie derivative can also be extended to compute the effect of one vector field on derivatives involving another; specifically, for vector fields f and g, the mixed Lie derivative is given by L_g L_f^{k-1} h(x) = \frac{\partial (L_f^{k-1} h(x))}{\partial x} g(x). This form is particularly useful for examining how inputs influence higher-order output derivatives in control systems. Geometrically, the Lie derivative L_f h(x) measures the directional change of the scalar field h along the integral curves (flow) of the vector field f, equivalent to the time derivative of h composed with the flow \phi(t, x) of f evaluated at t = 0, i.e., L_f h(x) = \frac{d}{dt} \bigg|_{t=0} h(\phi(t, x)). This interpretation is crucial for understanding output evolution in nonlinear systems, as it captures how observables vary under the system's dynamics without explicit time dependence. In the context of affine nonlinear models \dot{x} = f(x) + g(x)u, such derivatives facilitate the decomposition of system behavior into drift and control-influenced components. Key properties of the Lie derivative include its linearity in both arguments: L_{a f + b g} h = a L_f h + b L_g h and L_f (a h_1 + b h_2) = a L_f h_1 + b L_f h_2 for scalars a, b. It also satisfies a Leibniz rule akin to the chain rule: for smooth functions h_1 and h_2, L_f (h_1 h_2) = h_1 L_f h_2 + h_2 L_f h_1. For vector fields, the Lie derivative L_f g is the vector field whose i-th component is (L_f g)_i = \frac{\partial g_i}{\partial x} f, and it gives rise to the Lie bracket [f, g] = L_f g - L_g f, which quantifies the non-commutativity of flows generated by f and g. A brief derivation of the Lie bracket follows from considering the flows \phi_t^f and \phi_s^g: the commutator approximates the second-order term in the for , yielding [f, g](x) = \frac{\partial}{\partial s} \bigg|_{s=0} \left( \frac{\partial}{\partial t} \bigg|_{t=0} (\phi_t^f)_* g(\phi_s^g(x)) \right) - (f \leftrightarrow g), where (\cdot)_* denotes the pushforward, confirming [f, g] = L_f g - L_g f. These properties ensure the Lie derivative behaves consistently under coordinate transformations, making it invariant and suitable for geometric analysis in control.

Nonlinear System Models

Nonlinear dynamical systems relevant to feedback linearization are typically represented in state-space form as \dot{x} = f(x, u), where x \in \mathbb{R}^n denotes the state vector and u \in \mathbb{R}^m the input vector, with f: \mathbb{R}^n \times \mathbb{R}^m \to \mathbb{R}^n being a nonlinear function. For the analysis and application of feedback linearization, especially in single-input single-output (SISO) contexts, the systems are restricted to affine-in-control representations: \dot{x} = f(x) + g(x)u, where u \in \mathbb{R} is the scalar input, f: \mathbb{R}^n \to \mathbb{R}^n and g: \mathbb{R}^n \to \mathbb{R}^n are smooth vector fields, and the output is defined as y = h(x) with h: \mathbb{R}^n \to \mathbb{R} also smooth. This affine structure assumes linearity with respect to the control input, facilitating the design of feedback laws that cancel nonlinearities. Key assumptions underlying these models include the smoothness (at least C^\infty) of the vector fields f, g, and h over a domain containing the operating region, ensuring the existence of necessary derivatives and transformations. Additionally, basic controllability conditions are imposed, such as g(x) \neq 0 throughout the domain, to guarantee that the input influences the dynamics nontrivially. Feedback linearization distinguishes between input-output linearization, which focuses on rendering the input-to-output map linear to enable output tracking, and full state-space linearization, which transforms the entire internal dynamics into a linear controllable form. The former is particularly suited for systems where only outputs are measured or regulated, while the latter requires observability of the full state. Lie derivatives of the output function along the vector fields f and g are employed to compute the successive time derivatives of y, which are essential for assessing linearizability. Post-linearization, a diffeomorphic coordinate transformation \xi = \phi(x) can place the system into a normal form \dot{\xi} = A\xi + Bv, where A and B are matrices in controllable canonical form and v is a new linear input.

SISO Feedback Linearization

Relative Degree

In single-input single-output (SISO) nonlinear systems described by \dot{x} = f(x) + g(x)u and y = h(x), the relative degree r at a point x_0 is defined as the smallest integer r such that L_g L_f^k h(x_0) = 0 for all k = 0, 1, \dots, r-2, while L_g L_f^{r-1} h(x_0) \neq 0, where L_f and L_g denote the Lie derivatives along the vector fields f and g, respectively. This scalar relative degree quantifies the number of differentiations of the output y needed until the input u explicitly appears in the dynamics. The relative degree is computed iteratively through the Lie derivatives of the output function h(x). The first derivative is \dot{y} = L_f h(x), and subsequent derivatives follow as y^{(k)} = L_f^k h(x) for k < r-1, where the input does not appear since L_g L_f^k h(x) = 0. At the r-th derivative, the expression becomes y^{(r)} = L_f^r h(x) + L_g L_f^{r-1} h(x) \, u, with L_g L_f^{r-1} h(x) \neq 0, allowing the input to influence the output dynamics directly. This process relies on the chain rule for Lie derivatives, ensuring the computation reveals the input-output coupling structure. The relative degree is well-defined on an open domain containing x_0 if the conditions L_g L_f^k h(x) = 0 for k < r-1 and L_g L_f^{r-1} h(x) \neq 0 hold for all x in that domain, and r \leq n where n is the system dimension. In such cases, the system possesses a vector relative degree (r), meaning the differentials dh(x), dL_f h(x), \dots, dL_f^{r-1} h(x) are linearly independent, enabling a local coordinate transformation. The relative degree determines the dimension of the linearizable subsystem in Brunovsky canonical form, where the transformed coordinates yield a chain of r integrators. If r = n, the entire state space is linearizable via static state feedback, achieving full input-state linearization without internal dynamics.

Feedback Design and Coordinate Transformation

Once the relative degree r of the single-input single-output (SISO) nonlinear system \dot{x} = f(x) + g(x)u, y = h(x) has been determined, the feedback design proceeds by constructing a state feedback law and a diffeomorphism (coordinate transformation) that linearizes the input-output map. The coordinate transformation defines new coordinates \xi = (\xi_1, \dots, \xi_r)^T where \xi_i = L_f^{i-1} h(x) for i = 1, \dots, r, with L_f denoting the Lie derivative along the vector field f. The remaining n - r coordinates are given by \eta = \psi(x), where \psi is a smooth function such that the transformation (\xi, \eta) = \Phi(x) is a local diffeomorphism on a neighborhood of the operating point. Under this transformation, the feedback law is u = -\frac{L_f^r h(x)}{L_g L_f^{r-1} h(x)} + \frac{v}{L_g L_f^{r-1} h(x)}, where v is the new linear input and L_g L_f^{r-1} h(x) is the Lie derivative along g of the (r-1)-th Lie derivative of h along f. This law ensures that the r-th derivative of the output satisfies y^{(r)} = v. In the new coordinates, the dynamics of the \xi-subsystem become fully linear: \dot{\xi} = A \xi + b v, where A = \begin{pmatrix} 0 & 1 & 0 & \cdots & 0 \\ 0 & 0 & 1 & \cdots & 0 \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ 0 & 0 & 0 & \cdots & 1 \\ 0 & 0 & 0 & \cdots & 0 \end{pmatrix}, \quad b = \begin{pmatrix} 0 \\ 0 \\ \vdots \\ 0 \\ 1 \end{pmatrix}. This is the , which is controllable and allows standard linear control techniques to be applied to achieve desired input-output behavior. The \eta-subsystem evolves independently as \dot{\eta} = q(\xi, \eta). The feedback law and transformation are well-defined provided the invertibility condition \beta(x) = L_g L_f^{r-1} h(x) \neq 0 holds in the domain of interest, ensuring the denominator is nonzero and the transformation is invertible. Linearization is local if this condition and the diffeomorphism hold only in a neighborhood of the operating point; it is global if they hold over the entire state space, though global cases are rarer due to practical constraints on the functions involved.

Zero Dynamics

In the context of single-input single-output (SISO) feedback linearization, zero dynamics describe the internal behavior of the nonlinear system that persists after the output and its derivatives up to the relative degree are constrained to zero. Following a suitable coordinate transformation that separates the system into controllable and internal parts, the dynamics take the form \dot{\xi} = A \xi + B v for the linearizable output-related coordinates \xi and \dot{\eta} = q(\eta, \xi) for the internal coordinates \eta, where v is the new input. The zero dynamics are then obtained by restricting these internal dynamics to the case where \xi = 0, yielding \dot{\eta} = q_0(\eta) := q(\eta, 0). This restriction captures the unobservable evolution of the system when the output tracking is perfectly enforced. The stability of these zero dynamics plays a crucial role in determining the overall stabilizability of the system via output feedback. If the equilibrium \eta = 0 of \dot{\eta} = q_0(\eta) is unstable—such as when the linearized version around the equilibrium has eigenvalues with positive real parts—the internal states may diverge even as the output is successfully tracked to zero, rendering the full system unstabilizable by linearizing feedback alone. This phenomenon mirrors unstable transmission zeros in linear systems and highlights the limitations of input-output linearization for non-minimum-phase nonlinearities. To compute the zero dynamics, one first identifies the zero dynamics manifold as the largest invariant set contained in the level sets h(x) = 0, \dots, L_f^{r-1} h(x) = 0, where h(x) is the output function, L_f denotes the Lie derivative along the drift vector field f, and r is the relative degree. The dynamics are then restricted to this manifold by selecting the feedback input that maintains invariance, effectively projecting the vector fields f and g onto the tangent space of the manifold. This process, rooted in , allows explicit construction of q_0(\eta) in coordinates adapted to the manifold. A system is said to satisfy the minimum-phase condition if its zero dynamics are asymptotically stable at \eta = 0. This property ensures that the internal dynamics do not destabilize the closed-loop behavior, enabling robust output regulation and stabilization through high-gain feedback on the linearized subsystem. Seminal analyses show that minimum-phase SISO systems can achieve semiglobal asymptotic stability via such designs, provided the zero dynamics equilibrium is globally attractive.

MIMO Feedback Linearization

Multivariable Relative Degree

In the context of multi-input multi-output (MIMO) nonlinear systems, the relative degree from the single-input single-output (SISO) case is extended to a vector relative degree to characterize the input-output behavior and assess the potential for linearization. For a square MIMO system described by \dot{x} = f(x) + \sum_{j=1}^p g_j(x) u_j, y = h(x) \in \mathbb{R}^p, where x \in \mathbb{R}^n, the vector relative degree (r_1, \dots, r_p) is defined locally around a point x_0 such that, for each output h_i, the Lie derivatives satisfy L_{g_j} L_f^k h_i(x) = 0 for all j = 1, \dots, p and all k = 0, \dots, r_i - 2, while the row vector [L_{g_1} L_f^{r_i - 1} h_i(x), \dots, L_{g_p} L_f^{r_i - 1} h_i(x)] \neq 0 at x_0. This ensures that the input u first influences the r_i-th derivative of y_i through the vector fields g_j. The total relative degree is given by r = \sum_{i=1}^p r_i, which represents the dimension of the controllable subspace achievable via input-output linearization; if r = n, the system admits full state linearization provided additional structural conditions hold, such as the invertibility of the associated decoupling matrix. The decoupling matrix A(x) \in \mathbb{R}^{p \times p} is formed by the rows [a_{i1}, \dots, a_{ip}] where a_{ij} = L_{g_j} L_f^{r_i - 1} h_i(x), and its rank equals p at points where the vector relative degree is well-defined, ensuring the outputs can be independently influenced by the inputs. For the vector relative degree to exist, the output functions h_1, \dots, h_p must be such that their differentials dh_1, \dots, dh_p are linearly independent in a neighborhood of x_0, guaranteeing that the outputs are functionally independent. Computation of the vector relative degree involves iterative evaluation of multivariable Lie derivatives: starting from y = h(x), successive Lie derivatives L_f^k h(x) are calculated until the first order r_i - 1 where L_g L_f^{r_i - 1} h_i(x) \neq 0, with L_g denoting the matrix whose columns are L_{g_j}. This process highlights the number of differentiations needed to reveal input dependence for each output channel.

Decoupling Matrix and Feedback Law

In multivariable feedback linearization, the decoupling matrix plays a central role in achieving independent control channels for multi-input multi-output (MIMO) nonlinear systems of the form \dot{x} = f(x) + \sum_{j=1}^p g_j(x) u_j, y = h(x), where h(x) = \mathrm{col}(h_1(x), \dots, h_p(x)) and the system has a well-defined vector relative degree (r_1, \dots, r_p) with \sum r_i \leq n. The decoupling matrix A(x) is defined as the p \times p matrix whose i-th row and j-th column entry is the Lie derivative A_{ij}(x) = L_{g_j} L_f^{r_i - 1} h_i(x). This matrix captures the influence of each input u_j on the highest-order derivative of each output y_i before direct input appearance. For feedback linearization to proceed, A(x) must be invertible over a domain containing the operating region, ensuring uniform relative degree and the absence of singular configurations. The r_i-th time derivative of the output y_i is then given by y_i^{(r_i)} = L_f^{r_i} h_i(x) + \sum_{j=1}^p L_{g_j} L_f^{r_i - 1} h_i(x) u_j, or in vector form, \begin{bmatrix} y_1^{(r_1)} \\ \vdots \\ y_p^{(r_p)} \end{bmatrix} = b(x) + A(x) u, where b(x) = \begin{bmatrix} L_f^{r_1} h_1(x) \\ \vdots \\ L_f^{r_p} h_p(x) \end{bmatrix}. The feedback law that linearizes and decouples the input-output map is the static state feedback u = -A^{-1}(x) b(x) + A^{-1}(x) v, where v = \begin{bmatrix} v_1 \\ \vdots \\ v_p \end{bmatrix} is a new input vector. Substituting this law yields the desired dynamics y_i^{(r_i)} = v_i for each i = 1, \dots, p, transforming the coupled nonlinear input-output behavior into p independent scalar channels. To realize this linear structure internally, a diffeomorphic coordinate transformation is employed: \xi_{i,j}(x) = L_f^{j-1} h_i(x) for j = 1, \dots, r_i and i = 1, \dots, p, forming chains of length r_i for each output, with complementary coordinates \eta(x) spanning the remaining n - \sum r_i dimensions, assuming the required distributions are integrable. In these coordinates, the \xi-subsystem decouples into p linear chains: for each i, \dot{\xi}_{i,1} = \xi_{i,2}, \dots, \dot{\xi}_{i,r_i-1} = \xi_{i,r_i}, \dot{\xi}_{i,r_i} = v_i, or compactly, \dot{\xi}_i = A_i \xi_i + b_i v_i, where A_i is the r_i \times r_i companion (shift) matrix and b_i is the corresponding input vector (e.g., [0, \dots, 0, 1]^T). This allows independent linear control design for each chain, such as pole placement via v_i = -k_i^T \xi_i + w_i for reference tracking.

Internal Dynamics

In multivariable feedback linearization, the internal dynamics represent the unobservable subsystem remaining after the input-output linearization transformation, capturing the behavior constrained by the outputs being held at zero along with their derivatives up to the relative degrees. Specifically, for a system with outputs y_i = h_i(x) and vector relative degrees (r_1, \dots, r_p), the internal dynamics are defined on the manifold where y_i = \dot{y}_i = \dots = y_i^{(r_i-1)} = 0 for all i = 1, \dots, p. In the normal coordinates (\xi, \eta), where \xi encompasses the observable states and \eta the internal states, these dynamics are governed by \dot{\eta} = q(\eta, \xi), and restricting to the zero-output constraint \xi = 0 yields the zero dynamics \dot{\eta} = q(\eta, 0). The zero dynamics manifold is the submanifold of the state space invariant under the system dynamics and tangent to the maximal controlled invariant distribution contained in the kernel of the output map; it is characterized by the codistribution generated by the differentials dh_i, dL_f h_i, \dots, dL_f^{r_i-1} h_i for each output i, ensuring the constraints are integrable. This manifold generalizes the SISO zeroing manifold to handle coupled multivariable constraints, often requiring the for integrability verification. Stability of the internal dynamics is crucial for overall system performance, particularly for output regulation. The zero dynamics are hyperbolic if the Jacobian of q(\eta, 0) at the equilibrium has no eigenvalues on the imaginary axis; asymptotic stability implies the system is minimum phase, enabling robust asymptotic output tracking via the linearized subsystem. Conversely, unstable zero dynamics render the system non-minimum phase, leading to potential instability or bounded-but-not-convergent tracking errors, as the internal states may diverge despite stabilized outputs. When the total relative degree r = \sum_{i=1}^p r_i < n, where n is the system order, the internal dynamics have dimension n - r > 0, introducing complexity beyond the SISO case due to the higher-dimensional manifold and potential non-integrability issues in the multivariable codistribution. This dimensionality gap necessitates careful analysis of the zero dynamics' attractivity and invariance properties to ensure the constrained trajectories remain bounded. The decoupling matrix from the feedback law ensures input-output linearization without altering these internal dynamics, focusing control solely on the part.

Extensions

Input-Output Linearization

Input-output linearization is a technique in theory that focuses on rendering the input-output map of a linear through state , without necessarily linearizing the entire state dynamics. For a single-input single-output (SISO) nonlinear system described by \dot{x} = f(x) + g(x)u and y = h(x), where x \in \mathbb{R}^n, u \in \mathbb{R}, and y \in \mathbb{R}, the method applies when the output function h(x) has a well-defined relative r (with $1 \leq r \leq n) at points in a domain D_0. This relative r is the smallest integer such that the Lie derivative L_g L_f^{r-1} h(x) \neq 0, while L_g L_f^{i} h(x) = 0 for $0 \leq i < r-1. The approach targets the transformation y^{(r)} = v, where v is a new virtual input, thereby achieving a linear relationship between v and the output derivatives up to order r, while the remaining n - r internal dynamics are left unaddressed. The procedure involves differentiating the output y repeatedly until the input u explicitly appears, which occurs at the r-th : y^{(r)} = L_f^r h(x) + L_g L_f^{r-1} h(x) \, u. A static law is then designed as u = \alpha(x) + \beta(x) v, with \alpha(x) = -L_f^r h(x) / L_g L_f^{r-1} h(x) and \beta(x) = 1 / L_g L_f^{r-1} h(x), assuming \beta(x) is nonsingular in D_0. This yields the desired relation y^{(r)} = v, transforming the input-output behavior into a chain of r integrators: \dot{\xi}_i = \xi_{i+1} for i = 1, \dots, r-1, \dot{\xi}_r = v, and y = \xi_1, where \xi = \Phi(x) is a partial coordinate transformation. To achieve output tracking, a linear controller is applied to the virtual input v, such as v = y_d^{(r)} - \sum_{i=0}^{r-1} k_i (y^{(i)} - y_d^{(i)}), where y_d is the reference and k_i are gains ensuring of the resulting linear . When r < n, the system decomposes into these external (linearized) and internal \dot{\eta} = q(\eta, \xi), which must be analyzed separately for overall but are not transformed. Unlike full state feedback linearization, which requires the system to be invertible with relative degree n and a global diffeomorphism to linearize the entire state space, input-output linearization does not demand a full coordinate transformation on the state manifold. It is particularly useful for systems where the relative degree is less than the state dimension, allowing partial linearization even if the distribution \Delta = \operatorname{span}\{g, \text{ad}_f g, \dots, \text{ad}_f^{n-2} g\} does not span the tangent space, thus handling cases where full linearization is impossible. This partial approach simplifies design by focusing solely on output regulation, making it applicable to underactuated or partially observable systems without requiring invertibility of the full dynamics. A representative example is a mechanical system like a two-link robotic manipulator, modeled as M(q) \ddot{q} + C(q, \dot{q}) \dot{q} + G(q) = \tau, with output y = q (joint angles) and input \tau (torques). Here, the relative degree is r = 2 (since the first derivative involves velocities and the second brings in torques), and the feedback \tau = M(q) v + C(q, \dot{q}) \dot{q} + G(q) linearizes the input-output map to \ddot{q} = v. A linear controller on v, such as v = \ddot{q}_d - k_d (\dot{q} - \dot{q}_d) - k_p (q - q_d), enables precise tracking for the outputs, while internal dynamics related to unmodeled or higher-order effects remain nonlinear. This is advantageous in scenarios with known output measurements but unmodeled internal states, such as flexible structures in .

Robustness Considerations

Feedback linearization relies on exact knowledge of the system dynamics, particularly the nonlinear functions f(x) and g(x), to achieve precise cancellation of nonlinearities through state feedback. However, in practice, modeling errors or uncertainties in these functions result in incomplete cancellation, leaving residual nonlinearities that can degrade performance or cause instability. To address these sensitivities, robust variants have been developed, including adaptive feedback linearization, which estimates and adjusts uncertain parameters online to maintain linearization accuracy. For instance, adaptive schemes update controller gains based on real-time error signals, ensuring stability and tracking under parametric variations. Another approach integrates sliding mode control with feedback linearization to provide disturbance rejection; the sliding mode component enforces robustness against matched uncertainties and external perturbations by driving the system states onto a sliding surface, while linearization handles the nominal dynamics. Key limitations arise from potential singularities in the control design, such as when the scalar \beta(x) vanishes or the decoupling matrix A(x) becomes ill-conditioned, leading to unbounded control efforts or loss of . These issues can be mitigated in part by high-gain observers, which estimate unmeasured states robustly for output implementations, allowing even when full state information is unavailable. Post-2000 developments have focused on hybrid methods to enhance semi-global stability, such as combining feedback linearization with for systems not fully linearizable or with unmodeled dynamics, where recursively stabilizes subsystems. Similarly, integration with (MPC) on the linearized model optimizes trajectories while accounting for constraints and uncertainties, improving overall robustness. More recent advancements as of 2025 include data-driven feedback linearization using derivatives and stacked for systems with unknown dynamics, as well as learning-enhanced robust frameworks that incorporate adaptive mechanisms and to compensate for disturbances without precise models. Fixed-structure robust extensions have also been proposed to handle residual nonlinearities via additional terms. Computational challenges, like real-time evaluation of derivatives, are addressed through approximations or reduced-order models to enable practical .

References

  1. [1]
    [PDF] Lecture 13: Feedback Linearization
    Using control authority to transform nonlinear models into linear ones is one of the most commonly used ideas of practical nonlinear control design. Generally, ...
  2. [2]
    [PDF] Feedback Linearizing Control of Affine Nonlinear Systems
    Feedback linearizing controllers, essentially, use feedback laws that ”cancel” the nonlinear dynamics of the original plant and replacing them with a ”linear ...
  3. [3]
    [PDF] Feedback Linearization Of Nonlinear Systems
    The Problem of Feedback Linearization​​ A basic issue in control theory is how to use feedback in order to modify the original internal dynamics of a controlled ...<|control11|><|separator|>
  4. [4]
    Nonlinear Control Systems - SpringerLink
    In 1979, Alberto Isidori initiated a research program aimed at the extension of so-called "geometric theory" of multivariable linear systems, pioneered in the ...
  5. [5]
  6. [6]
    [PDF] Chapter 4 - Feedback Linearizing Control
    In this chapter, the basic theory of feedback linearization is presented and issues of particular relevance to process control applications are discussed.
  7. [7]
    Feedback Invariants for Nonlinear Systems - ScienceDirect
    This paper investigates finding invariants under nonlinear feedback, which describe the intrinsic nonlinearity of a control system that cannot be removed by  ...
  8. [8]
    (PDF) Feedback linearization - ResearchGate
    Aug 8, 2025 · This paper considers the problem of designing C1 (continuously differentiable) state feedback stabilizers for a class of 3-dimensional nonlinear ...
  9. [9]
    Linearization by output injection and nonlinear observers
    Observers can easily be constructed for those nonlinear systems which can be transformed into a linear system by change of state variables and output injection.
  10. [10]
    Partial and robust linearization by feedback - ResearchGate
    Aug 6, 2025 · The exact feedback linearization method was first proposed by Isidori et al. in the eighties [11] only for single-input singleoutput (SISO) ...
  11. [11]
    [PDF] The early days of geometric nonlinear control
    The paper (Brockett, 1973b) makes use of proper- ties of compact groups to develop a theory of systems evolving on spheres; this work is directly applicable to ...
  12. [12]
    Computation of multiple Lie derivatives by algorithmic differentiation
    Lie derivatives are often used in nonlinear control and system theory. In general, these Lie derivatives are computed symbolically using computer algebra ...Missing: seminal | Show results with:seminal
  13. [13]
    [PDF] Controllability and Lie Bracket - mimuw
    A nonlinear control system can be considered as a collection of dynamical systems (vector fields) parametrized by a parameter called control. It is natural to ...<|control11|><|separator|>
  14. [14]
  15. [15]
    [PDF] Chapter 2 Feedback linearization - Control Systems
    1. Put the system in companion form;. 2. Linearize the system through feedback linearization techniques;. 3. Control the system with standard linear methods ...
  16. [16]
    Feedback linearization - Scholarpedia
    Oct 25, 2011 · Feedback linearization is the process of determining a feedback law and a change of coordinates that transform a nonlinear system into a linear ...Missing: seminal | Show results with:seminal
  17. [17]
    Nonlinear Control Systems: An Introduction - SpringerLink
    The purpose of this book is to present a self-contained description of the fundamentals of the theory of nonlinear control systems.
  18. [18]
    [PDF] Lecture : Feedback Linearization
    Summary: This document follows the lectures on feedback linearization tought at the ... Definition 5 (Relative degree). The nonlinear system. ˙x = f(x) + g(x)u y ...
  19. [19]
    The zero dynamics of a nonlinear system - ScienceDirect.com
    The concept of zero dynamics of a nonlinear system was introduced about thirty years ago as nonlinear analogue of the concept of transmission zero of the ...
  20. [20]
    [PDF] Zero dynamics of nonlinear systems
    The main original motivation for introducing this concept was the ambition to develop systematic methods for asymptotic stabilization, with guaranteed region of ...
  21. [21]
    [PDF] Feedback Linearization March 9, 2022 1 Introduction 2 Examples f
    Mar 9, 2022 · Feedback linearization, in its simplest form, proceeds by differentiating the output mapping y = h(x) enough times so that all controls appear ...
  22. [22]
    [PDF] Nonlinear Systems and Control - Lecture 6 (Meetings 20-22 ...
    In the case of state feedback, i.e., ym = x, the controller takes the form u = γ(x, σ, e), where γ is designed such that there is a unique σss that satisfies.
  23. [23]
  24. [24]
    Adaptive Feedback Linearization: Stability and Robustness | IDEALS
    Author(s): Kanellakopoulos, Ioannis ; Issue Date: 1989-01 ; Keyword(s). Nonlinear systems; Adaptive regulation scheme; Stability; Robustness; Adaptive tracking.
  25. [25]
  26. [26]
    (PDF) Robust Feedback Linearization - ResearchGate
    Oct 7, 2014 · It is shown that a feedback linearized controller can be designed to be robust in face of parametric uncertainties, by simply pushing the ...Missing: considerations | Show results with:considerations
  27. [27]
  28. [28]
    A new approach to dynamic feedback linearization control of an ...
    Aug 6, 2025 · Three control laws with state feedback (linearizing feedback, High gain and backstepping) are proposed to address the problem of perfect ...
  29. [29]
    [PDF] A Feedback Linearized Model Predictive Control Strategy for Input ...
    May 2, 2024 · To this end, we develop a dual-mode Model Predictive. Control (MPC) solution starting from an input-output feedback linearized description of ...