Recap: LTI systems

This section contains a selection of the material from the module on discrete-time systems, which provides a brief recap of the general form and properties of linear time-invariant systems.

Linearity and time-invariance

A linear time-invariant system is a system having both the properties of linearity and time-invariance, which are described hereafter.

Linearity

A system is regarded to be linear if it is both additive and homogeneous. Suppose that we have a system, which we provide separately with two distinct input signals $x_1[n]$ and $x_2[n]$. These signals are individually transformed by the system into two output signals, $y_1[n]$ and $y_2[n]$ respectively. This can be represented intuitively as \begin{equation} x_1[n] \rightarrow y_1[n] \quad \text{and} \quad x_2[n] \rightarrow y_2[n]. \end{equation}

This system is called additive if and only if the output of the system driven by the sum of the distinct inputs equals the sum of the individual outputs. In other words, when the individual input signals are simultaneously applied to the system, the respective outputs are also observed simultaneously. This can be represented as \begin{equation} \text{Additive:} \quad x_1[n] + x_2[n] \rightarrow y_1[n] + y_2[n]. \end{equation}

Furthermore, this system is called homogeneous if and only if the output of this system produces a identically scaled output when driven by a scaled input. In other words, when the input signal is scaled with some scaling coefficient $c$, the output is also scaled with the same scaling coefficient. This can be represented as \begin{equation} \text{Homogeneous:} \quad c \cdot x_1[n] \rightarrow c \cdot y_1[n]. \end{equation}

When a system is both additive and homogeneous, the system is linear. This is represented as \begin{equation} \text{Linear:} \quad \alpha \cdot x_1[n] + \beta \cdot x_2[n] \rightarrow \alpha \cdot y_1[n] + \beta \cdot y_2[n], \end{equation} where $\alpha$ and $\beta$ are scaling coefficients.

Time-invariance

Another system property is time invariance. This property concerns signal delays or temporal shifts. A system is called time invariant if a temporal shift in the input signal results in the same temporal shift in the respective output signal. This can be represented as \begin{equation} \text{Time-invariance:} \quad x[n-n_0] \rightarrow y[n-n_0], \end{equation} where $n_0$ represents the temporal shift.

System architecture

A general LTI system can be represented using a so-called signal flow diagram as in Fig. 1. The system is driven by an input signal $x[n]$ and outputs the signal $y[n]$. The output is obtained by the addition of two main branches. The upper branch, which is connected to the input signal, transforms the input and delayed inputs into an output. This branch is also called the moving-average part of the system. The lower branch, which is connected to the output signal, transforms delayed outputs into a new output. This branch is also called the auto-regressive part of the system. A more detailed explanation of the individual building blocks can be found here.

Signal flow diagram of an LTI system.
Signal flow diagram of an LTI system.



Difference equation

The operations of the system can be represented mathematically by the weighted sum of the current input and delayed versions of the input and outputs. This mathematical description is called the difference equation. The output $y[n]$ at sample index $n$ can be calculated as \begin{equation}\label{eq:DE} y[n] = \underbrace{\sum_{k=0}^{M-1} b_k x[n-k]}_\text{moving-average} + \underbrace{\sum_{k=1}^{N-1} a_k y[n-k]}_\text{auto-regressive}, \end{equation} where two distinct terms can be distinguished. First is the moving-average part, which represents the upper branch of Fig. 1, and then follows the auto-regressive part, which represents the lower branch of Fig. 1. The weights are represented by the coefficients $a_k$ and $b_k$, respectively.

Impulse response

Besides the difference equation, there is another way of describing the functioning of a system. This representation is called the impulse response and will prove useful in the following sections. The impulse response is the output, or response, of a system when driven by a short impulse $\delta[n]$ and is denoted by $h[n]$. This can be described intuitively as \begin{equation} \delta[n] \rightarrow h[n], \end{equation} where $\delta[n]$ represents the Dirac delta pulse, which is defined as \begin{equation} \delta[n] = \begin{cases} 1, & \text{for } n=0, \newline 0, & \text{elsewhere}. \end{cases} \end{equation} Another way of representing the definition of the impulse response is \begin{equation} h[n] = y[n]\bigg\rvert_{x[n] = \delta[n]}, \end{equation} which simply states that the impulse response is the output of a system when the input of that system is given by a Dirac delta pulse.

The impulse response can be metaphorically compared with a tuning fork. A tuning fork is used to tune an instrument. The fork is hit against another object, after which it will resonate at a specific frequency. This frequency is used as a guideline for the tuning of an instrument. This can be compared to the discussion of the impulse response. The system (the tuning fork) is excited with an impulse (hitting it against a solid object) and will output a signal, also called the impulse response (the sound signal at the resonant frequency).

FIR and IIR filters

From the representation of the general LTI system, two different types of systems can be distinguished. These two types are finite impulse response (FIR) filters and infinite impulse response (IIR) filters. As the name indicates, the distinction is based on the length of the impulse response.

FIR filters are discussed in more detail here. FIR filters are the class of filters in which all auto-regressive coefficients $a_k$ are equal to zero. These filters effectively only contain the upper branch of the signal flow diagram in Fig. 1. When a FIR filter is excited by an impulse, this passes through the upper branch and is delayed a total of $M-1$ times; the delayed samples are weighted and summed to produced the output of the system. Mathematically, the impulse response of an FIR filter can be determined by substituting $\delta[n]$ for $x[n]$ in \eqref{eq:DE} and by setting all auto-regressive coefficients $a_k$ to zero as \begin{equation} \begin{split} h[n] &= \sum_{k=0}^{M-1} b_k \cdot \delta[n-k] + \sum_{k=1}^{N-1} 0 \cdot y[n-k], \newline &= \sum_{k=0}^{M-1} b_k \cdot \delta[n-k] = b_0\delta[n]+b_1\delta[n-1]+ …+ b_n\delta[n-M-1]. \end{split}\label{eq:impulse_FIR} \end{equation} The last step in this equation can be understood by noting that $\delta[n-k]$ only equals $1$ when $n=k$ holds. From this it can be noted that the length of the impulse response depends on the number of moving-average weights $b_k$. The number of moving-average weights $b_k$ depends on the length of the upper branch of the filter, which should be finite to allow for a practical implementation. Therefore the number of moving-average weights is also finite, leading to finite impulse response given by \eqref{eq:impulse_FIR}.

IIR filters on the other hand, contain non-zero auto-regressive coefficients. Intuitively one could understand the consequence of this by paying close attention to the created feedback loop in the system, corresponding to the lower branch in the signal flow diagram in Fig. 1. Suppose an impulse is applied to the input of an IIR filter. This signal will propagate through the moving-average branch (or directly, when only $b_0$ is non-zero) to the output. Since at least one of the auto-regressive coefficients is non-zero, this output will pass through the auto-regressive branch back to the output and the signal will therefore keep propagating through the feedback loop and will keep generating outputs. By applying a single impulse at the input of the system, the system will keep generating outputs, leading to an infinitely long impulse response.

In the following example we will derive the impulse response of a simple system with a non-zero auto-regressive coefficient.

Example


Calculate the analytical impulse response of a first-order auto-regressive filter.
A first-order auto-regressive filter is defined as a system where the moving-average coefficients are all zero, except for $b_0$, which equals $1$, and where all auto-regressive coefficients are zero, except for $a_1$. Without the coefficient $b_0$, no signal would propagate to the output of the system and therefore there would be no influence of the auto-regressive portion. This definition gives rise to the corresponding difference equation \begin{equation*} y[n] = x[n] + a_1 y[n-1]. \end{equation*}

Let us first calculate the first outputs of the system when driven by an input $x[n]$ to get some intuition of an auto-regressive process. If an arbitrary input signal $x[n]$ would be applied to the system, the first output of the system would be calculated as \begin{equation*} y[0] = x[0] + a_1 y[-1] = x[0] + a_1 0 = x[0], \end{equation*} where the signal output is assumed to be $0$ for $n < 0$ From this first output of the system, the second output can be determined as \begin{equation*} y[1] = x[1] + a_1 y[0] = x[1] + a_1 x[0] \end{equation*} and the third output as \begin{equation*} y[2] = x[2] + a_1 y[1] = x[2] + a_1 \left( x[1] + a_1 x[0] \right) = x[2] + a_1x[1] + a_1^2 x[0]. \end{equation*} Through these iterations the reader should be able to see some structure in the output signal. Using this structure an equation for the output $y[n]$ can be found as \begin{equation} y[n] = \begin{cases} \displaystyle \sum^{n}_{k=0} a_1^k x[n-k], &\text{for } n\geq 0, \newline
0, & \text{for } n < 0. \end{cases} \end{equation} By using the definition of the unit step function $u[n]$, this can also be written as \begin{equation*} y[n] = u[n]\sum^{n}_{k=0} a_1^k x[n-k]. \end{equation*}

Since the impulse response of the system is defined as the output of the system when driven by a Dirac delta pulse, i.e. $x[n] = \delta[n]$, the impulse response can be found as \begin{equation} h[n] = u[n] \cdot a_1^n = \begin{cases} a_1^n, &\text{for } n\geq 0, \newline 0, &\text{for } n<0. \end{cases} \end{equation} Based on the value of $a_1^n$ the impulse response will have a different shape. The figure below shows several examples of possible impulse response for different values of $a_1^n$.

Different impulse responses for a first-order auto-regressive filter, depending on the value of the auto-regressive coefficient $a_1$.
Different impulse responses for a first-order auto-regressive filter, depending on the value of the auto-regressive coefficient $a_1$.

System invertibility

A system $H(z)$ is invertible if the output $y[n] (n=(-\infty, \infty))$ can be used to uniquely define the input $x[n] (n=(-\infty, \infty))$. This is only possible if every unique value in $x[n]$ maps to a unique value in $y[n]$. This is called one-to-one mapping. Obtaining the inverse for any system is very difficult. However, if the system is linear time-invariant with an impulse response $h[n]$, the inverse $h_{inv}[n]$ can be defined as follows

\begin{equation*} \begin{split} x[n] \ast h[n] &= y[n]\newline y[n] \ast h_{inv}[n] &= x[n]\newline x[n] &= (x[n]\ast h[n])\ast h_{inv}[n] \end{split} \end{equation*} This can be simplified by utilizing the z-transform. \begin{equation*} \begin{split} X(z) H(z) &= Y(z)\newline Y(z) H_{inv}(z) &= X(z)\newline X(z) &= X(z) H(z) H_{inv}(z)\newline H_{inv}(z) &= \frac{1}{H(z)} \end{split} \end{equation*} If the systems consists of poles and zeros, than it can be split into its denominator and numerator. \begin{equation*} \begin{split} H(z) = \frac{N(z)}{D(z)}\newline H_{inv}(z) = \frac{D(z)}{N(z)} \end{split} \end{equation*}

Thus, the poles of the system become zeros in its inverse and vice versa. However the inverse is not always uniquely defined for every system, as shown by the following example.

Take a look at the relative pages for a refresher on the Z-transform and system function.

Example

Consider a system with the impulse response $h[n] = \delta[n]-\frac{1}{4}\delta[n-1]$. By performing the z-transform we end up with $H(z)=1-\frac{1}{4}z^{-1}$. Therefore, the inverse is equal to $H_{inv}(z)= \frac{1}{1-\frac{1}{4}z^{-1}}$, which has one pole at $z=\frac{1}{4}$.

If we choose the region of convergence (ROC) as $|z| > \frac{1}{4}$, the inverse system is causal and stable, and

\begin{equation*} h_{inv}[n] = (\frac{1}{4})^n u[n] \end{equation*} However, if we choose the ROC as $|z| < \frac{1}{4}$, the inverse system is noncausal and unstable \begin{equation*} h_{inv}[n] = -(\frac{1}{4})^n u[-n-1] \end{equation*}

All-pass filter

An all-pass filter let's through all frequencies with the same magnification factor, that is the magnitude of the frequency response is constant $$ | H_{ap}(e^{j\theta})| = 1 \hspace{3mm} \forall \theta $$ Constraining the system gain to be unitary implies that the poles and zeros of a rational system function to occur in mirrored pairs. Thus if $H(z)$ has a pole $z=\alpha_k$, $H(z)$ must also have a zero at the mirrored location $z=1/\alpha^\ast_k$. \begin{eqnarray*} \text{Complex } h[n] &:& H_{ap}(z) = \prod_{k=1}^{p} \frac{z^{-1}-\alpha_k^\ast}{1 - \alpha_k z^{-1}} \newline \text{Real } h[n] &:& H_{ap}(z) = \prod_{k=1}^{N_s} \frac{|\alpha_{k}|^2 - 2 \Re e \{ \alpha_k \} z^{-1}+ z^{-2}}{1 - 2 \Re e \{ \alpha_k \} z^{-1}+ |\alpha_{k}|^2 z^{-2}} \end{eqnarray*} The poles of a stable and causal all-pass filter $H(z)$ lie inside the unit circle, that is all $|\alpha_k|<1$. If the impulse response $h[n]$ is real-valued, the complex roots occur in conjugate pairs, and these conjugate pairs can be combined to form second-order factors from which all coefficients are real, as shown in the equation above.

Example


Show that the following system is all pass: $$ H(e^{j\theta})=\frac{\frac{1}{2} - e^{-j\theta} + e^{-j2 \theta}}{1 - e^{-j\theta} + \frac{1}{2} e^{-j2\theta}} $$ Give the pole zero plot and a rough sketch of the magnitude and phase response plots.
By replacing $e^{-j\theta}$ with the complex variable $z^{-1}$ we obtain the following system function: $$ H(z)= \frac{\frac{1}{2} - z^{-1} + z^{-2}}{1 - z^{-1} + \frac{1}{2} z^{-2}} $$ Because of the special structure of this equation, the absolute value results in: \begin{eqnarray*} |H(z)| &=&\sqrt{H(z) \cdot (H(z))^*}= \sqrt{H(z) \cdot H^*(z^{-1})}\\ &=& \sqrt{ \left ( \frac{\frac{1}{2} - z^{-1} + z^{-2}}{1 - z^{-1} + \frac{1}{2} z^{-2}} \right ) \cdot \left ( \frac{\frac{1}{2} - z + z^{2}}{1 - z + \frac{1}{2} z^{2}} \right )} \\ &=& \sqrt{\left ( \frac{z^{-2} \left ( \frac{1}{2}z^2 - z + 1 \right )}{z^{-2} \left ( z^2 - z + \frac{1}{2} \right )} \right ) \cdot \left ( \frac{\frac{1}{2} - z + z^{2}}{1 - z + \frac{1}{2} z^{2}} \right )} = 1 \\ \Rightarrow & & |H(e^{j\theta})| = \vert H(z)\vert_{|z|=1} =1 \end{eqnarray*} The pole and zeros are: $$ H(z) = \frac{(\frac{1}{2} \sqrt{2}- e^{j\frac{\pi}{4}}z^{-1}) (\frac{1}{2} \sqrt{2}- e^{-j\frac{\pi}{4}}z^{-1})}{(1- \frac{1}{2} \sqrt{2} e^{j\frac{\pi}{4}}z^{-1})(1- \frac{1}{2} \sqrt{2} e^{j\frac{\pi}{4}}z^{-1})} $$ as shown at the left hand side of the following plot:
A sketch of the magnitude and phase response is depicted at the right hand side of the figure.

Minimum phase systems

The simple example above illustrates that the knowledge of the impulse response of a LTI system does not uniquely specify its inverse. Additional information such as causality and stability would be helpful in many cases. This leads us to the concept of minimum-phase systems.

A minimum phase system is a system in which the system and its inverse are both stable and causal. To understand what this means we first need to know the conditions for stability and causality.

  • In a stable system the ROC contains the unit circle.
  • In a causal system the ROC is defined as going outward from a circle, with a radius larger than the largest pole in the system (excluding poles at infinity).
  • In a anti-causal system the ROC is defined as going inward from a circle, with a radius smaller than the smallest pole (excluding poles at zero).

Therefore, if a systems is minimum phase, its poles and zeros have the following properties. The ROC extends outwards from a circle with a radius larger than the largest pole, since the system needs to be causal. The system also needs to be stable. Therefore, the ROC includes the unit circle. Thus, the largest pole, and therefore all poles, need to be within the unit circle. The zeros need to be within the unit circle as well, since these rules also apply to the inverse.

A stable and causal LTI system has all its poles inside the unit circle. The zeros, however may lie anywhere in the $z$-plane. In order to obtain a minimum phase system, it is necessary to constraint the system $H(z)$ so that its inverse $G(z)=1/H(z)$ is also stable and causal. This requires that the zeros of $H(z)$ lie also inside the unit circle.

In other words, a stable and causal filter that has a stable and causal inverse is said to have minimum phase and so a minimum phase system has all of its poles and zeros inside the unit circle, as shown in Fig. 2.

Pole-zero plot and magnitude- and phase-response of minimum phase system.
The magnitude and phase response of a minimum phase system with three zeros inside the unit circle is depicted in Fig. 2.

When mirroring all zeros and poles of a stable and causal minimum phase system we obtain a maximum phase system; thus, a maximum phase system has all of its poles and zeros outside the unit circle, as shown in Fig. 3.

Pole-zero plot and magnitude- and phase-response of maximum phase system.

Fig. 3. shows the magnitude and phase response of a maximum phase system with three zeros outside the unit circle as depicted in the pole-zero plot. These zeros are obtained by mirroring all three zeros of the previous minimum phase system.

  • Mirroring a zero with respect to the unit circle does not change the shape of the magnitude response. Thus both minimum and maximum phase systems have, besides a constant factor, the same shape of the magnitude response characteristic.
  • The phase of a minimum phase system has less variations compared to the phase variations of a maximum phase system.
  • The group delay is defined as: $$ \tau (\theta) = - \frac{\text{d} \varphi \{ H(e^{j\theta}) \}}{\text{d} \theta}, $$ The group delay of a minimum phase system is minimal.

Minimum phase and all-pass decomposition

With the knowledge that we have about minimum phase and all-pass systems, we are able to show that any causal pole-zero system with system function $H(z)$ (without poles or zeros on the unit circle) can be decomposed as the product of an all-pass and a minimum phase system.

Let $H(z)$ be a non-minimum phase system with one zero $z=\frac{1}{a}, |a|<1$, outside the unit circle and all other poles and zeros on the inside of the unit circle. To decompose the system the steps to follow are:

  1. Factorize out all poles and zeros outside of the unit circle to create a minimum phase and maximum phase system. For our example $H(z)$ can be rewritten as: \begin{equation*} H(z) = H_{min1}(z)(a-z^{-1}) = H_{min1}(z)H_{max}(z), \end{equation*} where $H_{1}$ is minimum phase.
    Pole-zero plot of H(z).
    Pole zero plot of $ H(z)$, decomposed as a cascade of a minimum phase and a maximum phase systems.
  2. Create all the conjugate reciprocals of all poles and zeros that were factorized out. However, you cannot just add poles and zeros to a system for free. You will also need to add a zero on the same spot for every pole you add and vice versa. In our example, this becomes \begin{equation*} H(z) = H_{min1}(z)(a-z^{-1}) \frac{1-a^\ast z^{-1}}{1-a^\ast z^{-1}} = H_{min1}(z)H_{mix}(z) \end{equation*}
    Pole-zero plot of H(z).
    Pole zero plot of $ H(z)$, decomposed as a cascade of a minimum phase and a mixed phase systems.
  3. Match all the factored out poles and zeros with their conjugate reciprocal to obtain the all-pass part, and add the remaining poles and zeros the minimum phase system part. Finally, we obtain \begin{equation*} \begin{split} H(z) &= [H_{min1}(z)(1-a^\ast z^{-1})]\frac{a-z^{-1}}{1-a^\ast z^{-1}} \\ H(z) &= H_{min}(z)H_{ap}(z) \end{split} \end{equation*}
    Pole-zero plot of H(z).
    Pole zero plot of $ H(z)$, decomposed as a cascade of a minimum phase and an all-pass systems.

In the following exercise, you can try for yourself to apply these steps and obtain a minimum-phase all-pass decomposition.

Exercise


Factor the following system function as the product of a minimum phase system and an all-pass system: $$ H(z) = \frac{1 - 2 z^{-1}}{1 - 0.9 z^{-1}} $$
This can be shown as follows: \begin{eqnarray*} H(z) &=& \frac{1 - \color{green}{2} z^{-1}}{1 - 0.9 z^{-1}} \cdot \frac{1 - \color{red}{\frac{1}{2} z^{-1}}}{1 - \color{blue}{\frac{1}{2} z^{-1}}} = \frac{1 - \color{red}{\frac{1}{2}} z^{-1}}{1 - 0.9 z^{-1}} \cdot \frac{1 - \color{green}{2} z^{-1}}{1 - \color{blue}{\frac{1}{2}} z^{-1}} \newline &=& H_{min}(z) \cdot H_{ap}(z) \end{eqnarray*} First create a zero by mirroring the zero, which results in a new zero at $z=0.5$. This new zero has to be compensated by a new pole at the same position $z=0.5$. Then rearrange all old and new poles and zeros of $H(z)$ into two factors. The first factor contains the old pole and the new zero, which are both inside the unit circle and so this factor is minimum phase. The second factor contains the old zero and new pole which are each others mirrored versions and so this factor is all pass.
Exercise factorization.
These steps can be applied to any filter $H(z)$