upgrade
upgrade

🎛️Control Theory

Key Concepts in Time Domain Analysis

Study smarter with Fiveable

Get study guides, practice questions, and cheatsheets for all your subjects. Join 500,000+ students with a 96% pass rate.

Get Started

Why This Matters

Time domain analysis is the foundation of understanding how control systems actually behave—not in abstract frequency plots, but in real, measurable time. When you're tested on control theory, you're being asked to predict what happens when someone flips a switch, changes a setpoint, or disturbs a system. Can you determine if the response will be fast or sluggish? Will it overshoot dangerously or settle smoothly? Will there be persistent error? These questions drive everything from autopilot design to industrial process control.

The concepts here connect directly to system stability, transient performance, and steady-state accuracy—the three pillars examiners love to probe. You'll see these ideas appear in transfer function analysis, controller tuning problems, and system design questions. Don't just memorize definitions—know what each parameter tells you about system behavior and how changing one affects the others. That's what separates students who can solve problems from those who just recognize terms.


System Characterization Methods

These foundational techniques reveal how a system behaves by examining its response to standardized inputs. The key insight: simple test signals (steps, impulses, ramps) expose complex system dynamics in predictable ways.

Step Response Analysis

  • Most common test signal in control systems—evaluates how a system responds to a sudden, sustained change in input (like flipping a switch from 0 to 1)
  • Reveals stability, speed, and accuracy simultaneously—one test captures rise time, settling time, overshoot, and steady-state error
  • Foundation for controller tuning—PID parameters are often adjusted based on step response characteristics

Impulse Response Analysis

  • Mathematically fundamental—the impulse response h(t)h(t) completely characterizes any linear time-invariant (LTI) system
  • Transfer function connection—taking the Laplace transform of the impulse response directly yields H(s)H(s)
  • Convolution relationship—output for any input can be found by convolving that input with the impulse response

Ramp Response

  • Tests tracking ability—analyzes how well a system follows a continuously increasing input r(t)=tr(t) = t
  • Reveals velocity error—systems that perfectly track step inputs may still show persistent error for ramp inputs
  • Real-world relevance—models scenarios like following a moving target or maintaining speed on an incline

Compare: Step response vs. impulse response—both characterize LTI systems completely, but step response is easier to generate experimentally while impulse response connects directly to the transfer function. For FRQ problems asking you to "characterize system behavior," step response metrics are usually expected.


Transient Response Parameters

These metrics quantify how a system gets to its final value. Transient behavior reflects the interplay between system speed and stability—faster isn't always better if it causes oscillation.

Rise Time

  • Measures response speed—typically defined as time to go from 10% to 90% of final value (though 0% to 100% is sometimes used)
  • Inversely related to bandwidth—systems with higher bandwidth have shorter rise times
  • Trade-off with overshoot—reducing rise time often increases overshoot unless damping is carefully controlled

Settling Time

  • Defines practical stability—time required for response to stay within ±2% (or ±5%) of final value permanently
  • Dominated by dominant poles—for second-order systems, Ts4ζωnT_s \approx \frac{4}{\zeta \omega_n} for 2% criterion
  • Critical for system specifications—many real applications specify maximum allowable settling time

Overshoot and Undershoot

  • Overshoot quantified as percentage%OS=peak valuefinal valuefinal value×100%\%OS = \frac{\text{peak value} - \text{final value}}{\text{final value}} \times 100\%
  • Directly related to damping ratio—for second-order systems, %OS=eπζ/1ζ2×100%\%OS = e^{-\pi\zeta/\sqrt{1-\zeta^2}} \times 100\%
  • Undershoot indicates non-minimum phase behavior—systems with right-half-plane zeros exhibit initial response opposite to final direction

Transient Response

  • Umbrella term for pre-steady-state behavior—encompasses rise time, overshoot, settling time, and oscillation characteristics
  • Determined by pole locations—real poles give exponential responses; complex poles give oscillatory responses
  • Design target in most control problems—controllers are tuned to shape transient response to meet specifications

Compare: Rise time vs. settling time—rise time measures how quickly you reach the target, while settling time measures how quickly you stay there. A system can have fast rise time but long settling time if it oscillates. Exam questions often ask which parameter matters more for a given application.


Steady-State Performance

These concepts address what happens as tt \to \infty. A system can have perfect transient response but still fail if it doesn't accurately track the desired output.

Steady-State Error

  • Defined as ess=limte(t)e_{ss} = \lim_{t \to \infty} e(t)—the persistent difference between desired and actual output
  • Depends on system type and input type—Type 0 systems have finite error for step inputs; Type 1 systems have zero error for steps but finite error for ramps
  • Reduced by increasing loop gain—but higher gain can destabilize the system, creating a fundamental trade-off

Error Constants (Position, Velocity, Acceleration)

  • Position constant Kp=lims0G(s)K_p = \lim_{s \to 0} G(s)—determines step input error: ess=11+Kpe_{ss} = \frac{1}{1 + K_p}
  • Velocity constant Kv=lims0sG(s)K_v = \lim_{s \to 0} sG(s)—determines ramp input error: ess=1Kve_{ss} = \frac{1}{K_v}
  • Acceleration constant Ka=lims0s2G(s)K_a = \lim_{s \to 0} s^2G(s)—determines parabolic input error: ess=1Kae_{ss} = \frac{1}{K_a}

Compare: KpK_p vs. KvK_v vs. KaK_a—each constant corresponds to progressively more demanding inputs. A system with finite KpK_p but Kv=0K_v = 0 will track steps with some error but cannot track ramps at all. FRQs often give you a transfer function and ask you to calculate steady-state error for different input types.


Time Constants and Delays

These parameters capture how fast a system responds and what lags exist in the response. Time constants and delays fundamentally limit achievable closed-loop performance.

Time Constants

  • Defines exponential decay rate—for a first-order system G(s)=1τs+1G(s) = \frac{1}{\tau s + 1}, response reaches 63.2% of final value at t=τt = \tau
  • Relates to pole location—time constant τ=1p\tau = \frac{1}{|p|} where pp is the pole location
  • Multiple time constants in higher-order systems—dominant (slowest) time constant often determines overall response speed

Time Delay Effects

  • Modeled as esTde^{-sT_d} in transfer functions—represents pure transportation lag with no dynamics
  • Destabilizing influence—delays add phase lag without changing magnitude, reducing phase margin
  • Cannot be eliminated by feedback—unlike poles, delays cannot be "cancelled"; they fundamentally limit achievable bandwidth

Compare: Time constant vs. time delay—both slow system response, but time constants represent dynamics (energy storage) while delays represent pure lag (transport time). A time constant can be compensated with derivative action; a delay cannot. This distinction frequently appears in controller design problems.


Stability and System Structure

These analytical tools determine whether a system will behave predictably or diverge. Stability is non-negotiable—an unstable system is useless regardless of other performance metrics.

Stability Analysis in Time Domain

  • BIBO stability criterion—bounded input produces bounded output; requires all poles in left-half plane
  • Routh-Hurwitz criterion—algebraic test using characteristic polynomial coefficients; no need to find actual pole locations
  • Lyapunov methods—energy-based approach for nonlinear systems; if energy continuously decreases, system is stable

Pole-Zero Analysis

  • Poles determine natural response modes—each pole contributes an exponential or oscillatory term to the time response
  • Zeros affect response amplitude and shape—can cause overshoot reduction or non-minimum phase behavior
  • Dominant poles approximation—poles closest to imaginary axis dominate response; others can often be neglected

Root Locus Method

  • Tracks pole movement with gain—shows how closed-loop poles migrate as feedback gain KK varies from 0 to \infty
  • Visual stability assessment—system becomes unstable when locus crosses into right-half plane
  • Design tool for gain selection—choose KK to place poles at locations giving desired damping and speed

Compare: Routh-Hurwitz vs. root locus—Routh-Hurwitz tells you if a system is stable for a specific gain; root locus shows how stability changes across all gains. Use Routh-Hurwitz for quick stability checks; use root locus when designing for specific transient response.


Mathematical Representations

These frameworks provide the language for analyzing and designing control systems. State-space is more general; transfer functions are more intuitive for single-input, single-output systems.

State-Space Representation

  • Matrix formulation—system described by x˙=Ax+Bu\dot{x} = Ax + Bu and y=Cx+Duy = Cx + Du where xx is the state vector
  • Handles MIMO systems naturally—multiple inputs and outputs represented without partial fraction complications
  • Reveals internal dynamics—captures system behavior that may be hidden in transfer function (unobservable/uncontrollable modes)

Compare: Transfer function vs. state-space—transfer functions only capture input-output behavior; state-space reveals internal structure. A system can have identical transfer functions but different state-space realizations with different controllability properties. Modern control design (optimal control, observers) requires state-space formulation.


Quick Reference Table

ConceptBest Examples
Response speed metricsRise time, time constants, settling time
Stability indicatorsPole locations, Routh-Hurwitz, overshoot percentage
Steady-state accuracySteady-state error, KpK_p, KvK_v, KaK_a
System characterizationStep response, impulse response, ramp response
Graphical analysis toolsRoot locus, pole-zero plots
Mathematical modelsTransfer functions, state-space representation
Performance trade-offsRise time vs. overshoot, gain vs. stability
Practical limitationsTime delays, time constants

Self-Check Questions

  1. A second-order system has 15% overshoot and 2-second settling time. If you increase the damping ratio, what happens to both overshoot and rise time?

  2. Which error constant (KpK_p, KvK_v, or KaK_a) would you use to find steady-state error for a system tracking a linearly increasing reference? What system type is required for zero error?

  3. Compare the information provided by step response analysis versus impulse response analysis. Why might an engineer prefer one over the other in practice?

  4. A system has poles at s=1s = -1 and s=10s = -10. Which pole dominates the transient response, and approximately what is the system's time constant?

  5. Explain why time delay is more problematic for stability than a large time constant, even if both slow down the system response. How does each affect phase margin differently?