Fact-checked by Grok 2 weeks ago

Root locus analysis

Root locus analysis is a graphical method in for examining how the roots (poles) of the of a linear time-invariant vary with a system parameter, most commonly the open-loop gain K, which is varied from zero to infinity. This technique plots the trajectories of these closed-loop poles in the s-plane, providing insights into stability, , and performance characteristics such as ratio and . The root locus method was developed by American control theorist Walter R. Evans in the late 1940s while working at , where it addressed challenges in designing stable feedback systems for applications like autopilots and guidance systems during the post-World War II era of advancing aerospace technology. Evans first introduced the concept in his 1948 paper, where he described graphical methods for analyzing transient responses and pole movements in control systems, and also invented the Spirule, an analog slide-rule-like device to facilitate the computation of root loci. He expanded on this in 1950 with a detailed synthesis approach using root loci to design compensators and predict system behavior. Evans further popularized the technique through his 1954 book Control-System Dynamics, which formalized the rules and applications, making it a cornerstone of . The method's history was further chronicled in the 2025 book Into Stability: Walter R. Evans and the Story of Root Locus. As of November 2025. Key properties of the root locus include: the number of branches equals the number of open-loop poles; branches start at the open-loop poles (when K = 0) and terminate at the open-loop zeros or at (as K \to \infty); segments lie on to the left of an number of poles and zeros; and for systems with more poles than zeros, branches approach along asymptotes centered at \sigma = \frac{\sum \text{poles} - \sum \text{zeros}}{n - m} with angles \theta = \frac{(2q+1)180^\circ}{n-m} for q = 0, 1, \dots, n-m-1, where n and m are the numbers of finite poles and zeros, respectively. Additional rules cover departure/arrival angles at poles/zeros, breakaway/break-in points on , and crossings of the imaginary to assess margins. These rules enable sketching or computational plotting (e.g., via MATLAB's rlocus ) to evaluate how adjustments affect pole placement and thus . In practice, root locus analysis is essential for designing controllers in fields like , , and process control, as it visually reveals regions of (left-half s-plane) and allows selection of gain values for desired performance, such as minimal overshoot or rapid . Despite the rise of state-space and modern methods, it remains a fundamental tool for understanding parameter sensitivity and initial design iterations in linear systems.

Introduction

Definition and purpose

Root locus analysis is a graphical method in control systems engineering that depicts the trajectories of the closed-loop poles in the complex s-plane as a system parameter, typically the open-loop gain K, varies from 0 to infinity. This technique, developed by Walter R. Evans, allows engineers to visualize how the roots of the system's characteristic equation migrate, providing a qualitative understanding of dynamic behavior without exhaustive numerical computations. The paths, known as loci, originate at the open-loop poles (when K = 0) and terminate at the open-loop zeros (or infinity if there are fewer zeros than poles) as K increases. In a standard control system, the open-loop is denoted as G(s)H(s), where G(s) represents the and H(s) the path. The closed-loop poles are determined by solving the : $1 + K G(s) H(s) = 0. This equation defines the locations of the poles for any given K, and the root locus plots all such points satisfying the equation for K \geq 0. By examining the locus, one can identify regions where poles lie in the left-half s-plane for or cross into the right-half plane, indicating . The primary purpose of root locus analysis is to facilitate the design and tuning of feedback controllers by assessing margins, qualities (such as and overshoot), and steady-state accuracy without repeatedly solving the for various K values. It enables the selection of gain K that positions closed-loop poles at desired locations, such as those yielding specified ratios or natural frequencies, thereby optimizing system performance. For instance, if a pole location on the locus corresponds to acceptable , the associated K can be directly computed to achieve that configuration.

Historical development

The root locus method was pioneered by Walter R. Evans during the 1940s while working at in the aircraft industry, where he applied it initially to the design of systems during the post-World War II era of advancing technology and servo-mechanism development. Evans formalized the technique in his seminal 1948 paper, "Graphical Analysis of Systems," published in the Transactions of the , which introduced graphical methods for analyzing and in systems. Evans expanded upon this in his 1950 paper, introducing a detailed approach using root loci to compensators and predict system behavior. He further elaborated on these concepts in his 1954 book, Control-System Dynamics, published by McGraw-Hill, which provided a comprehensive treatment of root locus techniques and became a foundational text for engineers. The method gained prominence in the post-war era, influencing applications such as inertial guidance systems for missiles and like the F-86D, where it facilitated rapid of stable controllers essential for high-performance flight . By the 1970s, root locus analysis had been integrated into standard curricula, as evidenced by its detailed coverage in Katsuhiko Ogata's Modern Control Engineering (first edition, 1970), which popularized the among students and practitioners through systematic rules and examples. The to began in the 1960s, with early programs enabling automated plotting on computers like the IBM 7074, as described in technical reports that adapted Evans' graphical approach for numerical root locus generation. These developments laid the groundwork for later software tools, enhancing the 's accessibility beyond manual sketching.

Mathematical foundations

Characteristic equation

In control systems, the root locus analysis begins with the derived from the of a system. For a configuration, the is given by T(s) = \frac{K G(s)}{1 + K G(s)}, where K is the gain parameter and G(s) is the open-loop transfer function. The poles of this closed-loop system, which determine and , are the roots of the denominator set to zero, yielding the $1 + K G(s) = 0. This equation, introduced by Walter R. Evans in his foundational work on root locus methods, forms the basis for plotting how the closed-loop poles vary with K. For non-unity feedback systems, the general is T(s) = \frac{K G(s)}{1 + K G(s) H(s)}, where H(s) is the transfer function. The then becomes $1 + K G(s) H(s) = 0, or equivalently, when expressing the transfer function as L(s) = K \frac{N(s)}{D(s)} with N(s) and D(s) as the numerator and denominator polynomials, it simplifies to D(s) + K N(s) = 0. The open-loop poles are the roots of D(s) = 0, while the open-loop zeros are the roots of N(s) = 0. These poles and zeros define the starting and ending points of the root locus branches: as K = 0, the closed-loop poles coincide with the open-loop poles; as K \to \infty, the closed-loop poles approach the open-loop zeros or tend to along asymptotes if the number of poles exceeds the number of zeros. The number of branches in the root locus equals the of the , which is the degree of the D(s) + K N(s) = 0, corresponding to the number of open-loop . This ensures that each branch traces the migration of one closed-loop as K varies from 0 to \infty, providing insight into behavior without solving the equation for each K value.

Angle and magnitude conditions

In root locus analysis, the trajectory of the closed-loop poles as the gain K varies is determined by two fundamental conditions derived from the $1 + K G(s) H(s) = 0, or equivalently K G(s) H(s) = -1. These conditions ensure that for a point s in the , there exists a real value of K that places a closed-loop at that location. The angle condition specifies that s lies on the root locus if the phase of the open-loop transfer function satisfies \angle G(s) H(s) = (2k + 1) 180^\circ, where k is any . This arises because the argument of -1 is an multiple of $180^\circ, and for , the phase contribution from K is zero. Geometrically, this condition is interpreted using vectors in the s-plane: \angle G(s) H(s) equals the sum of the angles from all open-loop zeros to s minus the sum of the angles from all open-loop poles to s. A test point s satisfies the condition if the net is \pm 180^\circ, \pm 540^\circ, and so on. The magnitude condition complements the angle condition by determining the specific gain value: |K G(s) H(s)| = 1, or |G(s) H(s)| = 1/|K|. For a point s that already satisfies the angle condition, this equation yields the corresponding K > 0. In vector terms, |G(s) H(s)| is the product of the distances from s to all open-loop zeros divided by the product of the distances from s to all open-loop poles. Thus, K = \frac{\prod |s - z_i|}{\prod |s - p_j|}, where z_i are zeros and p_j are poles. Together, these conditions guarantee that s is a root of the characteristic equation for that K, tracing the locus as K varies from 0 to \infty. For the standard root locus assuming K > 0 (negative feedback), the angle condition uses odd multiples of $180^\circ. In contrast, for K < 0 (as in positive feedback or complementary root locus), the angle condition shifts to even multiples: \angle G(s) H(s) = 2k \cdot 180^\circ, while the magnitude condition becomes |K| \cdot |G(s) H(s)| = 1. This complementary locus, also introduced by Evans, explores pole locations for negative gains, often revealing additional stability insights but following similar vector-based construction principles.

Construction methods

Sketching rules

Root locus sketching rules offer a systematic, graphical approach to approximate the paths of closed-loop poles for varying gain K \geq 0 in linear feedback systems, facilitating manual construction without computational tools. Developed by in the mid-20th century, these heuristics stem from the angle and magnitude conditions of the characteristic equation $1 + K G(s)H(s) = 0, allowing rapid insights into stability and transient behavior. The rules apply to systems where the open-loop transfer function G(s)H(s) has poles at locations p_i (total n) and zeros at z_j (total m), with excess poles q = n - m. Consider the following standard rules for construction:
  • Number of branches: The root locus comprises exactly n branches, corresponding to the n open-loop poles of the system, as the closed-loop characteristic equation is of order n. Each branch represents the trajectory of one closed-loop pole as K varies.
  • Starting and ending points: All branches originate at the open-loop poles when K = 0 and terminate either at the open-loop zeros when K \to \infty or at infinity along asymptotes if q > 0. This reflects the continuity of pole locations from the open-loop to high-gain configurations.
  • Real-axis segments: Portions of the root locus lie on the real axis to the left of an odd number of real open-loop poles and zeros (counting multiplicities). Complex conjugate poles and zeros do not contribute to this count, as their loci are off-axis. To identify these segments, test points on the real axis by summing the number of poles and zeros to the right.
  • Asymptotes: When q > 0, the q branches approaching align with straight-line asymptotes. These asymptotes intersect the real axis at the \sigma = \frac{\sum \operatorname{Re}(p_i) - \sum \operatorname{Re}(z_j)}{q} and extend at \theta_k = \frac{(2k+1)180^\circ}{q} for k = 0, 1, \dots, q-1. This configuration captures the large-|s| behavior dominated by excess poles.
  • Symmetry: The entire root locus is symmetric about the real axis, ensuring that if a point s lies on the locus, its complex conjugate \bar{s} does too. This property arises because the characteristic equation has real coefficients, pairing complex poles and zeros.
Additional guidelines refine the sketch: For complex poles without nearby zeros, branches often depart at approximately $180^\circ relative to the real axis, while arrival angles at complex zeros follow complementary directions based on the angle condition. The locus intersects the imaginary axis at gains where the system is marginally stable, identifiable by the phase reaching -180^\circ. These rules collectively enable a qualitative yet informative plot, emphasizing conceptual system dynamics over precise coordinates.

Asymptotes and breakaway points

In root locus analysis, asymptotes describe the behavior of the locus branches as the K approaches , particularly when the number of finite open-loop poles exceeds the number of finite open-loop zeros. This occurs because excess branches must tend toward in the s-plane to satisfy the angle and magnitude conditions of the $1 + K G(s) H(s) = 0. The number of asymptotes equals p - m, where p is the number of finite poles and m is the number of finite zeros. The asymptotes intersect the real at the \sigma, calculated as \sigma = \frac{\sum \operatorname{Re}(p_i) - \sum \operatorname{Re}(z_j)}{p - m}, where \operatorname{Re}(p_i) and \operatorname{Re}(z_j) are the real parts of the finite and zero locations, respectively. This acts as the origin from which the asymptotes radiate. The angles of the asymptotes relative to the positive real are given by \phi_k = \frac{(2k + 1) 180^\circ}{p - m}, \quad k = 0, 1, \dots, p - m - 1. These angles ensure the total phase contribution aligns with the 180° condition for points on the locus at large |s|. For large values of K, the locus branches asymptotically approach these lines, providing insight into the high-gain and of the closed-loop system. Breakaway points occur on the root locus where multiple closed-loop poles coincide and then diverge, on the real axis, marking a transition from real to roots. These points lie on segments of the real axis between poles or zeros and are found by solving for locations where the gain K is stationary with respect to s, derived from K = -1 / G(s) H(s). Differentiating yields \frac{dK}{ds} = 0, which, letting G(s) H(s) = N(s)/D(s), leads to the condition N(s) D'(s) - N'(s) D(s) = 0. This results in a equation whose real between relevant poles and zeros indicate breakaway points. At these points, the locus changes, often corresponding to maximum or minimum K along the real axis. Break-in points are analogous to breakaway points but describe locations where complex conjugate branches reconverge onto the real axis, typically to the left of a zero or between poles. They are identified using the same differentiation method \frac{dK}{ds} = 0, with solutions selected based on the direction of branch entry onto the real axis, ensuring consistency with the overall locus topology. These points help refine sketches by pinpointing where the locus re-enters the real axis after departing as complex pairs. The jω-axis crossings, or points where the root locus intersects the imaginary axis, indicate the boundary between stable and unstable regions, corresponding to with sustained oscillations. These are located by applying the Routh-Hurwitz criterion to the , forcing a row of zeros in the Routh array to find the critical gain K and solving the auxiliary polynomial for the crossing frequency \omega. Alternatively, substitute s = j\omega into the and equate real and imaginary parts to zero, yielding equations for \omega and K. This method determines the maximum stable gain before poles enter the right-half plane.

Computational approaches

Numerical plotting techniques

Numerical plotting techniques enable the precise computation of root loci by solving the underlying mathematical conditions iteratively, providing greater accuracy than manual sketching for complex systems. The original approach, developed by Walter R. Evans, involved an iterative grid search in the complex s-plane to identify points where the open-loop satisfies the angle condition—namely, the argument of G(s) equals an odd multiple of 180 degrees—and the magnitude condition |G(s)| = 1/K for varying gain K. This method systematically evaluates candidate points on a predefined grid, computing phase contributions from poles and zeros to locate segments of the locus, though it was computationally intensive even in its early adaptations to analog or early computers. Modern algorithms improve upon this by parameterizing the root locus as continuous curves in the (s, K) parameter space, often using or predictor-corrector schemes to trace branches from open-loop poles toward zeros or . For instance, a predictor-corrector method initializes at known starting points (such as open-loop poles) and advances along the locus using approximations for prediction, followed by corrections to enforce both angle and magnitude conditions simultaneously, with adaptive step lengths to maintain convergence. These techniques solve for s along radial rays from the or predefined contours, iteratively adjusting to satisfy the conditions while handling multiple branches efficiently. In state-space representations, the locus corresponds to the eigenvalues of the closed-loop matrix A - k B C as k varies; numerical computation involves solving the generalized eigenvalue problem for the matrix pencil (sI - A, k B C) at discrete k values or using parametric eigenvalue solvers to trace the paths, which is particularly useful for multivariable systems where forms are cumbersome. To integrate root-finding capabilities, algorithms often discretize the gain over a range and compute the roots of the 1 + G(s) = 0 using robust polynomial solvers, such as those based on the eigenvalue decomposition or the Jenkins-Traub , yielding closed-loop poles at each that can be connected to form the locus. is frequently employed here, iterating on the complex variable s to minimize the residual of the , starting from initial guesses along predicted paths. For complex systems with high-order dynamics or multiple branches, numerical instability arises near breakaway points where loci converge or diverge, leading to ill-conditioned Jacobians in iterative solvers; this is mitigated by incorporating branching point detection—solving for multiple roots of the derivative of the —and employing higher-order corrections or variable-precision arithmetic to ensure smooth tracing without divergence. The output of these techniques is typically a parametric plot in the , displaying the real part of against the imaginary part for continuously varying K, often augmented with gain annotations and stability margins to facilitate analysis. This computational framework allows for high-fidelity , essential for verifying manual sketches and optimizing controller parameters in practical designs.

Software implementation and example

Root locus plots can be generated using specialized functions in popular control systems software packages, facilitating practical analysis without manual sketching. In MATLAB, the rlocus function from the Control System Toolbox computes and plots the root locus for a single-input single-output (SISO) defined by its . Similarly, the Python Control Systems Library provides the root_locus_plot function, which calculates the locus by solving for roots of the $1 + k G(s) = 0 over a range of gains k and displays the plot using . For more integrated design workflows, MATLAB's environment supports root locus visualization through tools like SISO Design Tool or Tuner, where users can interactively adjust parameters and observe locus changes in real-time simulations. To implement a root locus analysis, first define the open-loop G(s)H(s). In , create a transfer function object using the tf command, specifying numerator and denominator coefficients; for instance, for a system with no finite zeros and poles at specified locations, the denominator polynomial is entered as a vector. Then, invoke rlocus(sys) to generate the , which displays the locus branches starting from open-loop poles (at gain K=0) and includes markers indicating closed-loop pole locations for discrete gain values. In , use TransferFunction from the control module to define the , followed by root_locus_plot(sys), which produces an interactive with gain contours. Interpretation involves examining the for key features: branches migrate from poles toward asymptotes as K increases, and markers allow selection of gains yielding desired pole damping or natural frequency via tools like rlocfind in . These numerical methods underlie the software, discretizing the gain range and solving eigenvalue problems for root locations. Consider the example open-loop transfer function G(s) = \frac{1}{s(s+1)(s+2)}, with poles at s = 0, s = -1, and s = -2. The corresponding root locus features three branches originating at these poles and extending to , as there are no finite zeros. One branch starts at the pole at s = -2 and extends leftward along the real axis to -∞. The other two branches start at s = -1 and s = 0, move toward each other along the real axis between -1 and 0, meet at a breakaway point near s ≈ -0.42, and then diverge into the as a pair of complex conjugates. The asymptotes, determined by the excess of poles over zeros (three), emanate from the at s = -1 with angles of $60^\circ, $180^\circ, and -60^\circ. The following pseudocode outlines the implementation in MATLAB:
sys = tf([1], [1 3 2 0]);
rlocus(sys);
This generates the plot, where the locus crosses the imaginary axis at \pm j \sqrt{2} for K = 6, marking the onset of marginal stability; for K > 6, branches enter the right-half plane, indicating instability. To assess damping, select points on the locus (e.g., via interactive tools) where the angle from the negative real axis yields a desired damping ratio \zeta; for instance, dominant complex poles near the $45^\circ line correspond to \zeta \approx 0.707, guiding gain selection for oscillatory performance. In Python, the equivalent code is:
import control as ct
import matplotlib.pyplot as plt
sys = ct.tf([1], [1, 3, 2, 0])
ct.root_locus_plot(sys)
plt.show()
This approach enables rapid iteration in controller design, confirming stability margins and characteristics for the .

Applications

Stability and performance analysis

Root locus analysis serves as a fundamental tool for evaluating the of linear time-invariant systems by visualizing the trajectories of closed-loop poles as the K varies from 0 to \infty. A exhibits absolute if all branches of the root locus lie entirely within the left-half of the s-plane for the relevant range of K, ensuring that all closed-loop poles have negative real parts and thus produce bounded outputs for bounded inputs. If any portion of the locus enters the right-half plane, the becomes unstable for those gain values, as poles with positive real parts lead to exponentially growing responses. This graphical assessment provides a direct and intuitive means to determine the range of K for which the remains , distinct from frequency-domain methods like the , which assesses via of the critical [point -1](/page/Point_No._1) + j0 but does not explicitly plot pole movements. Beyond , root locus plots enable detailed performance analysis by relating closed-loop locations to key metrics, particularly for systems dominated by pairs. The natural frequency \omega_n, which governs the speed of the oscillatory response, is determined by the magnitude of the dominant : \omega_n = |s| where s is the location. The damping ratio \zeta, indicative of the decay rate and overshoot, is given by \zeta = -\frac{\Re(s)}{|s|} with \Re(s) denoting the real part of s. Poles closer to the imaginary (smaller |\Re(s)|) yield lower , increasing oscillatory tendencies, while the distance to the imaginary correlates with margin, offering a measure of relative . For instance, a margin is larger when poles are farther left, providing robustness against gain perturbations. These -derived parameters allow engineers to predict and tune dynamic behavior without full time-domain simulation. Performance specifications such as percentage overshoot and settling time are estimated directly from dominant pole positions on the root locus, facilitating gain selection for desired responses. The approximate percentage overshoot for a second-order approximation is \%OS \approx 100 \exp\left( -\frac{\zeta \pi}{\sqrt{1 - \zeta^2}} \right), enabling selection of K to achieve targets like \zeta \approx 0.6 for roughly 10% overshoot, where the locus intersects lines of constant damping (rays at angle \cos^{-1} \zeta from the negative real axis). Settling time to within 2% of the final value is estimated as t_s \approx 4 / |\Re(s)|, emphasizing the need for sufficiently negative real parts to ensure rapid decay. By identifying points on the locus that satisfy both \zeta and \omega_n requirements—such as placing dominant poles at s = -\zeta \omega_n \pm j \omega_n \sqrt{1 - \zeta^2}—an optimal K is computed using the magnitude condition K = 1 / |G(s)H(s)| at that point, balancing responsiveness and damping. Despite its utility, root locus analysis is inherently limited to linear time-invariant systems, assuming constant parameters and neglecting nonlinearities, unmodeled dynamics, or time-varying effects that could alter actual and . It relies on second-order approximations for transient predictions, which hold only if nondominant poles are at least five times farther left than dominant ones and zeros do not significantly influence the response; otherwise, higher-order effects may invalidate estimates. These constraints highlight the method's role as a preliminary rather than a comprehensive solution for complex or real-world systems.

Controller design principles

Root locus analysis serves as a foundational tool in controller design by visualizing how closed-loop pole locations vary with , enabling engineers to select parameters that achieve desired and specifications such as damping ratio, , and steady-state error. In the design process, the root locus of the uncompensated plant is first sketched to identify limitations in achievable pole placements, after which compensators are added to reshape the locus iteratively until it passes through regions corresponding to the target dominant poles. This approach leverages the angle and magnitude conditions to ensure the selected satisfies the while meeting criteria like overshoot and . Lead and lag compensators are commonly designed using root locus to modify the open-loop pole-zero configuration and thus alter the trajectory of the locus. A lead compensator, typically of the form G_c(s) = K_c \frac{s + z}{s + p} with |z| < |p|, introduces a zero closer to the than its , pulling branches of the root locus to the left in the s-plane to increase damping and improve transient response speed. Conversely, a compensator, where |z| > |p|, places the nearer the to boost low-frequency and reduce steady-state with minimal impact on high-frequency transients, as the added pole-zero pair forms a that slightly shifts the locus rightward. Lead- combinations integrate both by first applying the lead portion for enhancement and then the lag for reduction, ensuring the locus aligns with specifications like a constant K_v \geq 20. Proportional-derivative (PD) and proportional-integral (PI) controllers represent simplified dynamic compensators whose effects on the root locus facilitate targeted performance improvements. A PD controller, G_c(s) = K_p + K_d s = K_d (s + \frac{K_p}{K_d}), adds a zero near the origin that attracts locus branches leftward, enhancing and reducing overshoot in systems with insufficient . In contrast, a PI controller, G_c(s) = K_p + \frac{K_i}{s} = \frac{K_i (s + \frac{K_p}{K_i})}{s}, introduces a pole at the origin and a zero nearby, extending low-frequency branches to eliminate steady-state error for step inputs by increasing the system type, though it may slow convergence if not tuned carefully. The iterative design process begins with analyzing the uncompensated 's locus to pinpoint deficiencies, such as branches entering the right half-plane or failing to achieve desired , then adds compensator poles and zeros to redirect the locus toward the specified region in the . For instance, in gain scheduling for nonlinear or varying operating conditions, root locus insights guide the selection of gain values at multiple points, varying K dynamically to maintain across regimes, as parameterized in networked PI implementations. A representative case involves improving a second-order like G(s) = \frac{1}{s(s+1)}, where the dominant at s = 0 leads to a slow response; a controller places a zero at approximately s = -3.6 to reshape the locus, yielding dominant poles at approximately s = -3.1 \pm j4.2 for a damping ratio \zeta \approx 0.59 and ≈1.5 seconds.

Extensions

Discrete-time systems (z-plane)

In discrete-time systems, the root locus technique is adapted to the z-plane, where the characteristic equation takes the form $1 + z^{-1} G(z) H(z) = 0, with G(z) representing the discrete plant and H(z) the , incorporating the inherent one-sample delay from digital computation and . Alternatively, for systems expressed via difference equations, it can be written as z^N + K \cdot \text{num}(z) = 0, where N is the system order and \text{num}(z) is the numerator scaled by K. This formulation allows of closed-loop pole locations as K varies from 0 to \infty. The root locus in the z-plane begins at the open-loop poles when K = 0, often including a pole at z = 1 for systems with steady-state tracking requirements, such as those incorporating an , and terminates at the open-loop zeros or extends to infinity if there are fewer zeros than poles. Unlike the s-plane, stability requires all closed-loop poles to lie inside the unit , defined by |z| < 1, ensuring bounded responses to bounded inputs. The construction rules mirror those in the continuous domain, with loci symmetric about the real axis and the number of branches equal to the system order, but interpretation focuses on encirclements of the unit disk rather than the imaginary axis. The angle and magnitude conditions for points on the locus are analogous, requiring \angle G(z) H(z) = \pm 180^\circ (2k + 1) for integer k, while the magnitude condition is |z^{-1} G(z) H(z)| = 1/K. This z-plane behavior arises from the mapping z = e^{sT}, where T is the sampling period, transforming s-plane stability regions into curved spirals within the unit disk. Key differences include potential circular trajectories of branches around the unit disk due to the exponential mapping, contrasting the linear asymptotes in the s-plane. To emulate continuous designs, the bilinear transform s = \frac{2}{T} \frac{z-1}{z+1} warps the z-plane onto the s-plane, mapping the unit circle to the imaginary axis and facilitating direct application of s-domain tools like Routh-Hurwitz criteria. A representative example is a digital velocity control system with plant poles at z=0.5 and z=1 (including an integrator), compensated by adding a zero at z=0.2, yielding the open-loop transfer function G(z) = \frac{K (z-0.2)}{(z-0.5)(z-1)}. The root locus starts at the open-loop poles and curves inside the unit circle as K increases, with the added zero shifting the locus to allow stable pole placement. For instance, at K \approx 0.27, a damping ratio \zeta \approx 0.7 is achieved, with roots at a distance of approximately 0.66 from the origin, providing desired settling times while maintaining stability inside the unit disk.

Multivariable root locus

In multi-input multi-output (MIMO) systems, the root locus technique is generalized to analyze the migration of closed-loop poles as feedback gains vary, providing insights into stability and performance beyond the single-input single-output (SISO) case. The fundamental setup involves the open-loop transfer function matrix G(s) \in \mathbb{C}^{m \times p}, where m is the number of outputs and p the number of inputs, and a constant gain matrix K \in \mathbb{R}^{p \times m}. The characteristic equation defining the closed-loop poles is \det(I + K G(s)) = 0, where I is the identity matrix of appropriate dimension. This equation determines the values of the complex variable s for which the system has poles, and the root locus traces these poles as elements of K are scaled or varied. Equivalently, the loci correspond to the points s where the eigenvalues of -G(s) equal the negative reciprocals of the eigenvalues of K, or more precisely, the locus of the eigenvalues of -G(s)^{-1} as the gain parameter changes. The generalized root locus plots the trajectories of these closed-loop eigenvalues in the complex plane as a scalar gain k > 0 multiplies G(s) (i.e., K = k I) or as a full gain K varies along parameterized paths, such as constant singular values or structured updates. In Evans' form adapted for MIMO systems, the locus is characterized by solving \det(\lambda I + G(s)) = 0 for \lambda = 1/k, where the branches represent the values of s satisfying this generalized eigenvalue problem for varying \lambda. This formulation highlights the spectral nature of the analysis, treating the root locus as the preimage under G(s) of the negative real axis in the eigenvalue plane, analogous to the SISO condition G(s) = -1/k. Unlike SISO loci, which consist of n distinct branches for an n-th , MIMO loci can exhibit multiplicity and interdependence due to the structure. Key challenges in multivariable root locus analysis arise from the non-unique branching of loci, which may form closed loops or reside on multi-sheeted Riemann surfaces rather than simple curves, complicating manual sketching and numerical computation. Loop interactions further exacerbate this, as changes in one channel can influence distant poles through cross-coupling, leading to unanticipated eigenvalue migrations that affect overall system coordination. To mitigate these issues and enhance robust design, () is employed on the return difference I + K G(s), decomposing the influence into principal directions and magnitudes; this allows assessment of the minimum for margins and identification of sensitive directions for gain scheduling. Applications of the multivariable root locus are prominent in modern control, particularly for state feedback design, where the loci guide the selection of K to assign eigenvalues for specified damping ratios and natural frequencies in high-order systems. In (LQR) synthesis, the technique approximates optimal gain selection by tracing eigenvalue paths under quadratic cost variations, revealing trade-offs in state weighting without full Riccati solutions; for instance, it highlights how increased control effort shifts loci toward more stable regions while respecting limits. These tools are especially valuable in and process control, where interactions demand precise pole placement for robust performance.

References

  1. [1]
    Why is the Root Locus Plot Important? - Swarthmore College
    The root locus plot indicates how the closed loop poles of a system vary with a system parameter (typically a gain, K).Simple Example to Motivate... · Another Way of Looking at the...
  2. [2]
    Graphical Analysis of Control Systems | IEEE Journals & Magazine
    Graphical Analysis of Control Systems. Abstract: The purpose of this paper is to demonstrate some graphical methods for finding the transient response of a ...
  3. [3]
    Request Rejected
    - **Status**: Insufficient relevant content.
  4. [4]
    Control-system Dynamics - Walter R. Evans - Google Books
    Aest amplifier amplitude analysis approximately asymptotes attenuation block ... root-locus plot rotation servo shaft shown in Fig sine wave sinusoidal ...
  5. [5]
    Derivation of Root Locus Rules - Swarthmore College
    Root locus rules include symmetry, number of branches, starting/ending points, locus on real axis, asymptotes, break-away/in points, and angles of departure/ ...
  6. [6]
    [PDF] The Root Locus Method - MIT OpenCourseWare
    The root locus is the set of values s for which p‹q holds, and k is any positive real value. (For reasons that will become clear later, this is the definition ...
  7. [7]
    [PDF] Root Locus Procedure
    Oct 26, 2015 · The root locus method plots the trajectories of the roots of the characteristic equation when certain parameters in the system vary. Given the ...
  8. [8]
    [PDF] EE 370L Controls Laboratory Laboratory Exercise #7 Root Locus ...
    Thus, root locus is a family of curves of the roots of . A common technique in control system design is to plot the root locus corresponding to a plant in ...
  9. [9]
    Root Locus Plots In Control Systems - Electrical4U
    May 7, 2024 · Introduced in 1948 by Evans, the root locus technique helps represent any physical system using a transfer function.
  10. [10]
    [PDF] Systems Analysis and Control - Lecture 12: Root Locus
    Definition 1. The Root Locus of ˆG(s) is the set of all poles of k ˆG(s). 1 + ... Next Lecture: Constructing the Root Locus. M. Peet. Lecture 12: Control Systems.
  11. [11]
    Graphical Analysis of Control Systems - ADS
    Graphical Analysis of Control Systems. Evans, Walter R. Abstract. The purpose of this paper is to demonstrate some graphical methods for finding the transient ...
  12. [12]
    Graphical Analysis of Control Systems - Semantic Scholar
    Graphical Analysis of Control Systems · W. Evans · Published in Transactions of the American… 1948 · Engineering.
  13. [13]
    Evans - The Root Locus Paper | PDF | Guidance System - Scribd
    3) Evans' "root locus" method provided a new way to design stable feedback control systems, which was important for inertial guidance systems needed for missile ...Missing: original | Show results with:original
  14. [14]
    Root locus diagrams by digital computer
    Root locus diagrams by digital computer Adaptation of root locus method to programming on IBM 7074 digital computer ... August 1, 1966. Subject Category.Missing: history | Show results with:history
  15. [15]
    Root Locus: Example 3
    N(s)= s + 3, and D(s)= s2 - 1 s - 2. Characteristic Equation is 1+KG(s)H(s)=0, or 1+KN(s)/D(s)=0, or D(s)+KN(s) = s2 - 1 s - 2+ K( s + 3 ) = 0. Completed Root ...
  16. [16]
    10.3 Evans Root Locus Construction Rule # 1: Beginning, End and ...
    This rule deals with the beginning and end of the Root Locus plot and its symmetry. We begin by writing the closed loop system characteristic equation in the ...
  17. [17]
    [PDF] massachusetts institute of technology
    Nov 9, 2007 · H(s) = 1. Thus every pole on the root locus should satisfy 1 + KG(s) = 0. Because s is a complex number, KG(s) = −1 leads to two requirements:.
  18. [18]
    General Information
    Insufficient relevant content. The provided URL (https://ieeexplore.ieee.org/document/4050735) does not contain accessible text or specific details about angle and magnitude conditions for root locus. The page appears to be a general IEEE Xplore entry without viewable content based on the input provided.
  19. [19]
    [PDF] Root Locus Techniques I - Aerostudents
    Evans, W. R. Control System Synthesis by Root Locus Method. AIEE Transactions, vol. 69,. 1950, pp. 66-69. Evans, W. R. Graphical Analysis of Control Systems.
  20. [20]
    None
    ### Summary of Root Locus for Negative Gain K<0, Complementary Locus, and Angle Condition Differences
  21. [21]
    The Root Locus Rules - MIT
    All root locus rules can be directly traced to the characteristic equation, 1+L(s)=0. If we assume that the loop transfer function can be written as L(s)=KL ...Missing: derivation | Show results with:derivation
  22. [22]
    Rules for Making Root Locus Plots - Swarthmore College
    The table below summarizes how to sketch a root locus plot (K≥0). This is also available as a Word Document or PDF.Missing: Evans 1954
  23. [23]
    [PDF] 2.004 Dynamics and Control II - MIT OpenCourseWare
    A breakaway point is the point on a real axis segment of the root locus between two real poles where the two real closed-loop poles meet and diverge to become ...
  24. [24]
    10.6 Evans Root Locus Construction Rule # 4: Break-Away and ...
    Break-away points occur when closed loop poles meet and break away from the real axis. Break-in points occur when complex segments enter the real axis.
  25. [25]
    [PDF] Feedback Control Systems - Lecture – Chapter 8 – Root Locus ...
    Sep 10, 2013 · 8.2 Defining the root locus. 8.3 Properties of the root locus. 8.4 ... 8.9 Root locus for positive-feedback systems. 8.10 Pole sensitivity.
  26. [26]
    Application of numerical analysis to root locus design of feedback ...
    Root finding methods from numerical analysis enable the solution of the problem to be one of convergent iteration rather than trial and error.Missing: computing | Show results with:computing
  27. [27]
    [PDF] arXiv:2003.04672v1 [eess.SY] 10 Mar 2020
    Mar 10, 2020 · This paper presents a numerical method to plot the root-locus of SISO dead-time systems, calculating starting points, poles, and roots entering ...
  28. [28]
  29. [29]
  30. [30]
    rlocus - Root locus of dynamic system - MATLAB - MathWorks
    The root locus of a dynamic system contains the closed-loop pole trajectories as a function of the feedback gain k (assuming negative feedback). Root loci ...
  31. [31]
    control.root_locus_plot — Python Control Systems Library 0.10.2 ...
    root_locus_plot. Root locus plot. Calculate the root locus by finding the roots of 1 + k * G(s) where G is a linear system and k varies over a range of gains.
  32. [32]
    Root Locus Design - MATLAB & Simulink - MathWorks
    Root locus design is a common control system design technique in which you edit the compensator gain, poles, and zeros in the root locus diagram.
  33. [33]
    [PDF] Root locus analysis
    Root locus analysis. 22 / 41. Page 23. 23/41. Root locus. Example. Routh-Hurwitz criterion: s3 + 3s2 + 2s + k = 0. Routh array: s3. : 1. 2 s2. : 3 k s1. : (6 - ...
  34. [34]
    [PDF] Chapter Eight Root Locus Control Design 8.3 Common Dynamic ...
    Several common dynamic controllers appear very often in practice. They are known as PD, PI, PID, phase-lag, phase-lead, and phase-lag-lead controllers.
  35. [35]
    [PDF] Section 6: Design of Lead and Lag Controllers using Root Locus
    Principles of Lead design via. Root Locus. • The compensator adds poles and zeros to the P(s) in the root locus procedure.
  36. [36]
    Extras: Designing Lead and Lag Compensators
    A phase-lead compensator tends to shift the root locus toward to the left in the complex s-plane. This results in an improvement in the system's stability and ...
  37. [37]
    None
    ### Summary of Root Locus Design for Controllers
  38. [38]
    [PDF] On the Gain Scheduling for Networked PI Controller Over IP Network
    This parameterization enables PI gain scheduling to be tractable for real-time on-line analysis with existing theories such as root locus so that the control ...
  39. [39]
    None
    ### Summary of Discrete-Time Root Locus in z-Plane
  40. [40]
    None
    Below is a merged summary of the root locus analysis in the z-plane, combining all the information from the provided segments into a single, comprehensive response. To retain maximum detail and ensure clarity, I will use a table in CSV format where appropriate to organize the dense information (e.g., equations, examples, and key points), followed by a narrative summary that ties everything together. Since the system has a "no thinking token allowed" constraint, I will focus on directly synthesizing the content without additional interpretation or elaboration beyond what is provided.
  41. [41]
    None
    ### Summary of Discrete-Time Root Locus in z-Plane (Source: Lecture Note by Dr. Ahmad Al-Mahasneh)
  42. [42]
  43. [43]
    [PDF] STABILITY ANALYSIS TECHNIQUES - Dr. Gregory L. Plett
    A transform that satisfies these requirements is the bilinear transform. ... This is exactly the same form as the Laplace-transform root locus. Plot ...
  44. [44]
    [PDF] Design Example: Digital Control of a Velocity Servo: - MIT
    K(z-0.2). (z-0.5)(z-1) , which has the root locus: Now consider adding an integrator to get ess = 0 . This implies adding a pole at the origin in s or at unity ...
  45. [45]
  46. [46]
    Sketching rules for the multivariable root locus design technique
    Sketching rules for the multivariable root locus design technique ... Abstract: New results are summarized which extend the Evans (1950) root locus design ...
  47. [47]
    Multivariable Nyquist-Bode and multivariable Root-Locus techniques
    Multivariable Nyquist-Bode and multivariable Root-Locus techniques. Abstract: The theory of algebraic functions is used to extend to multivariable systems the ...