Feedback linearization

Summary

Feedback linearization is a common strategy employed in nonlinear control to control nonlinear systems. Feedback linearization techniques may be applied to nonlinear control systems of the form

Block diagram illustrating the feedback linearization of a nonlinear system
[1]

(1)

where is the state, are the inputs. The approach involves transforming a nonlinear control system into an equivalent linear control system through a change of variables and a suitable control input. In particular, one seeks a change of coordinates and control input so that the dynamics of in the coordinates take the form of a linear, controllable control system,

[2]

(2)

An outer-loop control strategy for the resulting linear control system can then be applied to achieve the control objective.

Feedback linearization of SISO systems edit

Here, consider the case of feedback linearization of a single-input single-output (SISO) system. Similar results can be extended to multiple-input multiple-output (MIMO) systems. In this case,   and  . The objective is to find a coordinate transformation   that transforms the system (1) into the so-called normal form which will reveal a feedback law of the form

 [3]

(3)

that will render a linear input–output map from the new input   to the output  . To ensure that the transformed system is an equivalent representation of the original system, the transformation must be a diffeomorphism. That is, the transformation must not only be invertible (i.e., bijective), but both the transformation and its inverse must be smooth so that differentiability in the original coordinate system is preserved in the new coordinate system. In practice, the transformation can be only locally diffeomorphic and the linearization results only hold in this smaller region.

Several tools are required to solve this problem.

Lie derivative edit

The goal of feedback linearization is to produce a transformed system whose states are the output   and its first   derivatives. To understand the structure of this target system, we use the Lie derivative. Consider the time derivative of (2), which can be computed using the chain rule,

 

Now we can define the Lie derivative of   along   as,

 

and similarly, the Lie derivative of   along   as,

 

With this new notation, we may express   as,

 

Note that the notation of Lie derivatives is convenient when we take multiple derivatives with respect to either the same vector field, or a different one. For example,

 

and

 

Relative degree edit

In our feedback linearized system made up of a state vector of the output   and its first   derivatives, we must understand how the input   enters the system. To do this, we introduce the notion of relative degree. Our system given by (1) and (2) is said to have relative degree   at a point   if,

  in a neighbourhood of   and all  
 

Considering this definition of relative degree in light of the expression of the time derivative of the output  , we can consider the relative degree of our system (1) and (2) to be the number of times we have to differentiate the output   before the input   appears explicitly. In an LTI system, the relative degree is the difference between the degree of the transfer function's denominator polynomial (i.e., number of poles) and the degree of its numerator polynomial (i.e., number of zeros).

Linearization by feedback edit

For the discussion that follows, we will assume that the relative degree of the system is  . In this case, after differentiating the output   times we have,

 

where the notation   indicates the  th derivative of  . Because we assumed the relative degree of the system is  , the Lie derivatives of the form   for   are all zero. That is, the input   has no direct contribution to any of the first  th derivatives.

The coordinate transformation   that puts the system into normal form comes from the first   derivatives. In particular,

 

transforms trajectories from the original   coordinate system into the new   coordinate system. So long as this transformation is a diffeomorphism, smooth trajectories in the original coordinate system will have unique counterparts in the   coordinate system that are also smooth. Those   trajectories will be described by the new system,

 

Hence, the feedback control law

 

renders a linear input–output map from   to  . The resulting linearized system

 

is a cascade of   integrators, and an outer-loop control   may be chosen using standard linear system methodology. In particular, a state-feedback control law of

 

where the state vector   is the output   and its first   derivatives, results in the LTI system

 

with,

 

So, with the appropriate choice of  , we can arbitrarily place the closed-loop poles of the linearized system.

Unstable zero dynamics edit

Feedback linearization can be accomplished with systems that have relative degree less than  . However, the normal form of the system will include zero dynamics (i.e., states that are not observable from the output of the system) that may be unstable. In practice, unstable dynamics may have deleterious effects on the system (e.g., it may be dangerous for internal states of the system to grow unbounded). These unobservable states may be controllable or at least stable, and so measures can be taken to ensure these states do not cause problems in practice. Minimum phase systems provide some insight on zero dynamics.

Feedback linearization of MIMO systems edit

Although NDI is not necessarily restricted to this type of system, lets consider a nonlinear MIMO system that is affine in input  , as is shown below.

 

(4)

It is assumed that the amount of inputs is the same as the amount of outputs. Lets say there are   inputs and outputs. Then   is an   matrix, where   are the vectors making up its columns. Furthermore,   and  . To use a similar derivation as for SISO, the system from Eq. 4 can be split up by isolating each  'th output  , as is shown in Eq. 5.

 

(5)

Similarly to SISO, it can be shown that up until the  ’th derivative of  , the term  . Here   refers to the relative degree of the  'th output. Analogously, this gives

 

(6)

Working this out the same way as SISO, one finds that defining a virtual input   such that

 

(7)

linearizes this  'th system. However, if  ,   can obviously not be solved given a value for  . However, setting up such an equation for all   outputs,  , results in   equations of the form shown in Eq. 7. Combining these equation results in a matrix equation, which generally allows solving for the input  , as is shown below.

 

(8)

See also edit

Further reading edit

  • A. Isidori, Nonlinear Control Systems, third edition, Springer Verlag, London, 1995.
  • H. K. Khalil, Nonlinear Systems, third edition, Prentice Hall, Upper Saddle River, New Jersey, 2002.
  • M. Vidyasagar, Nonlinear Systems Analysis, second edition, Prentice Hall, Englewood Cliffs, New Jersey, 1993.
  • B. Friedland, Advanced Control System Design, facsimile edition, Prentice Hall, Upper Saddle river, New Jersey, 1996.

References edit

  1. ^ Isidori, Alberto (1995). Nonlinear Control Systems (Third ed.). Springer-Verlag London. p. 5. ISBN 978-1-4471-3909-6.
  2. ^ H. Nijmeijer and A. van der Shaft, Nonlinear Dynamical Control Systems, Springer-Verlag, p. 163, 2016.
  3. ^ Isidori, Alberto (1995). Nonlinear Control Systems (Third ed.). Springer-Verlag London. p. 147. ISBN 978-1-4471-3909-6.

External links edit

  • Fabio Celani and Alberto Isidori, ed. (2009). "Feedback linearization". Scholarpedia. Retrieved 31 December 2022.
  • ECE 758: Modeling and Nonlinear Control of a Single-link Flexible Joint Manipulator – Gives explanation and an application of feedback linearization.
  • ECE 758: Ball-in-Tube Linearization Example – Trivial application of linearization for a system already in normal form (i.e., no coordinate transformation necessary).
  • Wolfram language functions to do feedback linearization, compute relative orders, and determine zero dynamics.