Fact-checked by Grok 2 weeks ago

Phugoid

The phugoid is a long-period, lightly damped oscillatory in the longitudinal dynamics of an , characterized by a slow interchange of kinetic and that results in gradual variations in and altitude while maintaining a nearly constant , with corresponding variations in pitch attitude. This motion typically manifests as a but persistent , with periods ranging from 30 seconds to over a minute depending on the aircraft's speed and configuration. The phugoid mode was first identified by British engineer Frederick W. Lanchester in 1897 through experiments with gliders, where he observed the oscillatory behavior in uncontrolled longitudinal flight. Lanchester coined the term "phugoid," derived from word phugein meaning "to flee," to describe the wandering, fugitive-like nature of the motion, though he intended it to evoke "to fly." In the early , mathematician G. H. Bryan further formalized its mathematical description in 1911, establishing it as one of the two primary longitudinal dynamic modes alongside the short-period oscillation. In terms of key characteristics, the phugoid's is low, typically around 0.1 to 0.5 rad/s (or approximately \sqrt{g/u_0}, where u_0 is the forward speed), leading to periods of about 46 seconds for large like the during approach. Its ratio is very light, often between 0.01 and 0.1 (proportional to the drag-to-lift ratio C_D/C_L), allowing the to persist for several cycles without significant decay unless corrected by pilot input or augmentation systems. Unlike the short-period mode, which involves rapid and angle-of-attack changes with heavy and periods under 10 seconds, the phugoid primarily affects velocity and flight-path angle, making it sensitive to factors like and variations. This mode is analyzed using linearized state-space equations or transfer functions in flight control design, where low can pose challenges for instrument flying and .

Fundamentals

Definition and Characteristics

The phugoid is a long-period, in an aircraft's longitudinal motion, involving a cyclic exchange between kinetic and that manifests as periodic variations in , altitude, and angle while maintaining a nearly constant . This mode arises from disturbances such as gusts or inputs under conditions, resulting in a sinusoidal flight path where the aircraft alternately climbs (gaining altitude and losing speed) and descends (gaining speed and losing altitude). Unlike more rapid dynamic responses, the phugoid emphasizes over immediate , often persisting for multiple cycles if unaddressed. Key characteristics of the phugoid include a typical period of 20 to 100 seconds, during which oscillations remain small, generally 1 to 5 degrees, accompanied by variations of 10 to 30% of the trim speed and altitude excursions of hundreds to thousands of feet. For instance, in high-speed flight tests of the , phugoid maneuvers exhibited periods around 137 to 151 seconds, with speed changes of ±0.02 and altitude variations of ±2,600 feet. These oscillations occur exclusively in the longitudinal plane, involving forward speed, flight path angle, and , with negligible coupling to roll or yaw motions. The mode is lightly damped, often requiring several minutes for significant energy dissipation due to the low influence of aerodynamic relative to lift. The phugoid is distinct from other ; it contrasts with the short-period mode, a high-frequency, heavily damped focused on rapid changes in and rate with little alteration in or altitude. In comparison to lateral-directional modes like the spiral (a slow, non-oscillatory bank angle ) and (a coupled yaw-roll ), the phugoid remains purely longitudinal without involving sideslip or banking tendencies. Pilots commonly observe the phugoid as "porpoising"—a gentle, undulating motion in and altitude during otherwise stable flight, often noticeable after releasing controls following an abrupt input.

Physical Mechanism

The phugoid oscillation arises from a periodic exchange between an aircraft's , associated with variations in forward speed, and , linked to changes in altitude, in the absence of external control inputs. This occurs as the aircraft's total remains approximately constant during unforced longitudinal motion, with and aerodynamic forces mediating the transfer. For instance, a speed increase leads to a net downward force component that converts into through altitude gain, and vice versa. Aerodynamic forces, particularly and , play a central role in driving this exchange, primarily through variations in the angle of attack that alter generation while contributes to gradual speed decay. As the angle of attack increases, rises to initiate a climb, reducing airspeed; conversely, a decrease allows descent and speed , with acting tangentially to the flight path to dissipate over cycles. In approximations for unpowered flight, is often treated as constant or zero, emphasizing the balance between , , and in the glide-like motion. In a typical free-flight with fixed controls, the begins with an initial disturbance that raises the angle of attack, increasing and causing the to climb while losing speed and avoiding through gradual energy transfer. This is followed by a , where reduced allows descent, converting back to as speed increases, completing one cycle of the undulating . The motion thus manifests as a long-period in speed and altitude, with the flight path curving sinusoidally over distances on the order of tens of miles. Aircraft parameters qualitatively influence the oscillation's amplitude and period: an aft center of gravity reduces longitudinal stability, potentially increasing amplitude by weakening restoring moments. These factors interact to shape the mode's behavior without direct control intervention. For large-amplitude phugoids, nonlinear effects such as stall at high angles of attack or compressibility impacts at transonic speeds cause deviations from ideal sinusoidal behavior, introducing asymmetries or rapid energy loss that linear models cannot capture.

Mathematical Modeling

Linearized Equations of Motion

The linearized for the phugoid mode are derived from the full nonlinear six-degrees-of-freedom (6-DOF) equations of aircraft dynamics by focusing on longitudinal perturbations in the vertical plane, assuming symmetric flight and neglecting lateral-directional motions. Key assumptions include small perturbations from a steady trimmed condition (typically level flight at constant speed), small angles for (e.g., \sin \theta \approx \theta, \cos \theta \approx 1), constant mass and , negligible variations, and an Earth-fixed reference frame with flat-Earth approximation. These simplifications reduce the system to a fourth-order model governing perturbations in forward (u), vertical (w), rate (q), and angle (\theta). The derivation begins with the nonlinear longitudinal equations in body axes, derived from Newton's laws applied to the aircraft's center of gravity: \dot{u} = \frac{X}{m} - g \sin \theta - q w + \dot{w}_e \dot{w} = \frac{Z}{m} + g \cos \theta + q u - \dot{u}_e \dot{q} = \frac{M}{I_{yy}} \dot{\theta} = q where X and Z are the axial and normal forces, M is the pitching moment, m is the aircraft mass, I_{yy} is the moment of inertia about the pitch axis, g is gravitational acceleration, and \dot{u}_e, \dot{w}_e account for Earth rotation (often neglected). Perturbations are introduced around the trim condition (denoted by subscript 0), such that total variables are u = u_0 + \Delta u, etc., and forces/moments are expanded in Taylor series: X = X_0 + \frac{\partial X}{\partial u} \Delta u + \frac{\partial X}{\partial w} \Delta w + \cdots. At trim, steady-state terms balance (\dot{u}_0 = 0, etc.), and higher-order terms are dropped for small perturbations, yielding the linearized form. The position equations (\dot{x} = u \cos \theta - w \sin \theta \approx u, \dot{z} = -u \sin \theta + w \cos \theta \approx - \theta u_0 + w) are often decoupled for dynamic analysis, focusing on the velocity and attitude states. The resulting linearized equations incorporate dimensional stability derivatives, such as X_u = \frac{\partial X}{\partial u}, Z_w = \frac{\partial Z}{\partial w}, and M_q = \frac{\partial M}{\partial q}, which capture aerodynamic sensitivities. These derivatives relate to nondimensional coefficients, for example, the lift curve slope C_{L\alpha} = \frac{\partial C_L}{\partial \alpha} (where \alpha \approx w / u_0 is the angle of attack), which influences Z_w \approx - \bar{q} S C_{L\alpha} with dynamic pressure \bar{q}, wing area S, and air density. Pitch damping M_q arises from tail and wing contributions, typically negative for stability, while gravity terms like -g \cos \theta_0 in the u-equation and g \sin \theta_0 in the w-equation introduce coupling between speed and altitude perturbations central to phugoid dynamics. In state-space formulation, the system is expressed as \dot{\mathbf{x}} = A \mathbf{x} + B \mathbf{u}, where the state vector is \mathbf{x} = [\Delta u, \Delta w, q, \Delta \theta]^T, \mathbf{u} includes control inputs (e.g., elevator deflection \delta_e), and A is the stability matrix. For level flight trim (\theta_0 = 0, \alpha_0 \approx 0), the matrix A simplifies to: A = \begin{bmatrix} \frac{X_u}{m} & \frac{X_w}{m} & 0 & -g \\ \frac{Z_u}{m} & \frac{Z_w}{m} & u_0 & 0 \\ \frac{M_u}{I_{yy}} & \frac{M_w}{I_{yy}} & \frac{M_q}{I_{yy}} & 0 \\ 0 & 0 & 1 & 0 \end{bmatrix} with elements adjusted for approximations like Z_q / m \approx u_0 (from kinematic relations). Nondimensionalization is often applied by scaling states (e.g., \hat{u} = \Delta u / u_0, \hat{\alpha} = \Delta w / u_0) and time (\tau = t \cdot u_0 / \bar{c}, with mean chord \bar{c}), yielding dimensionless derivatives like X_u' = X_u \cdot \bar{c} / u_0 for computational convenience and modal analysis. This form enables simulation of phugoid responses while preserving the essential coupling between speed, altitude, and pitch.

Mode Analysis and Eigenvalues

The mode analysis of the phugoid begins with the eigenvalue problem derived from the linearized longitudinal , formulated in state-space form as \dot{x} = A x, where A is the system matrix incorporating stability derivatives. The is obtained by solving \det(A - \lambda I) = 0, which yields a fourth-order and four eigenvalues representing the longitudinal modes: typically two pairs for the phugoid and short-period modes. The phugoid roots are the complex conjugate pair with a small negative real part \sigma (indicating light ) and low imaginary part \omega (corresponding to the ), expressed as \lambda = \sigma \pm i \omega, where typical values for are \sigma \approx -0.001 to -0.004 and \omega \approx 0.01 to $0.1 rad/s. For example, in a at 0.8 and 40,000 ft, the phugoid eigenvalues are \lambda = -0.0033 \pm 0.0672i. The period is calculated as T = 2\pi / \omega, yielding 30 to 100 seconds for commercial jets, while the time constant (e-folding time) is \tau = 1 / |\sigma|, often resulting in time to half-amplitude of approximately 170 to 250 seconds due to the light ratio \zeta = -\sigma / \sqrt{\sigma^2 + \omega^2} \approx 0.01 to $0.05. The eigenvectors associated with the phugoid eigenvalues define the mode shapes, revealing the relative amplitudes and phases of state perturbations such as forward speed u, \theta, \alpha, and q. In the phugoid mode, u and altitude h (derived from flight path angle integration) exhibit coupled oscillations that are effectively out of phase by 180°, with maximum speed coinciding with minimum altitude during descent; \theta shows a 180° phase shift relative to speed, such that nose-up attitude occurs during the low-speed climb phase, while \alpha and q remain small. For the Boeing 747 example, the normalized eigenvector components are approximately A_u = 0.036, A_\theta = 1, A_q = 0.0012, A_\alpha = 0.017, emphasizing the dominance of speed and . To simplify the full fourth-order , approximation methods eliminate the fast short-period , reducing the phugoid to a second-order model focused on speed and flight path angle. A common for the phugoid is \omega_p \approx \sqrt{2g / V_0}, where g is and V_0 is trim speed, providing a baseline independent of detailed but refined in advanced models to account for lift curve slope C_{L\alpha} and effects. More precise forms incorporate aerodynamic influences, such as \omega_p \approx \sqrt{(2g / V_0) \cdot (C_{L\alpha} / (2 - C_{D\alpha}))}, where C_{D\alpha} is the derivative with respect to , improving accuracy for conventional airplanes by including and contributions. These approximations facilitate preliminary design assessments without full eigenvalue computation.

Stability and Control

Damped and Undamped Behavior

In the ideal undamped phugoid, the real part of the eigenvalue (σ) is zero, resulting in perpetual oscillations that exchange kinetic and without decay, maintaining constant amplitude in speed and altitude. This theoretical case assumes no dissipative forces, leading to neutral where the neither returns to trim nor diverges. In real , the phugoid exhibits lightly damped behavior, typically with a ζ below 0.1, causing oscillations to decay exponentially over several cycles due to energy losses from parasite and propulsion effects such as variation with speed. The is proportional to the (C_D/C_L), which remains small in efficient designs, and increases with higher parasite or thrust lapse rates that reduce power output as speed rises. For example, in a at low , ζ ≈ 0.013, enveloping speed and altitude variations that contract gradually. The phugoid is inherently stable for most powered , characterized by a negative σ that ensures eventual return to , though is weak enough to require pilot or intervention over long periods. Rare neutral (σ ≈ 0) occurs in highly efficient gliders with large C_L/C_D ratios, approaching undamped oscillations, while unstable cases (positive σ) can arise in high-altitude flights where reduced air density diminishes . Environmental factors influence phugoid ; vertical gusts from excite the mode, amplifying initial oscillations, while alters stability—negative shear enhances , but positive shear can render the mode aperiodic and unstable for gradients exceeding certain thresholds. In hypersonic vehicles, the phugoid often diverges without control due to unique aerodynamic and propulsion interactions at extreme speeds. Simulations of phugoid responses reveal time histories where altitude and airspeed trace elliptical paths, with the envelope contracting over 4–10 cycles in damped cases, illustrating the slow energy dissipation that distinguishes it from faster modes.

Mitigation Strategies

Aircraft design significantly influences phugoid mitigation by optimizing parameters that enhance inherent damping and stability margins. The horizontal tail sizing, typically expressed through the tail volume coefficient, contributes to longitudinal static stability, which indirectly improves phugoid damping by increasing the restoring moments during speed-altitude exchanges. A forward-shifted center of gravity (CG) raises the static margin, thereby boosting natural phugoid damping ratios, while strict CG limits during certification prevent aft positions that could degrade mode stability. Fly-by-wire systems enable relaxed static stability configurations, where smaller tails reduce drag, and electronic augmentation compensates for any phugoid susceptibility without compromising overall handling. Pilot techniques emphasize proactive to interrupt the phugoid cycle before oscillations amplify. adjustments maintain constant , countering the kinetic-potential energy interchange, while subtle stick inputs apply to excursions without inducing short-period modes. Avoiding abrupt maneuvers, such as sudden changes, prevents phugoid , as the mode's long period (often 30-90 seconds) provides ample time for these interventions. These methods rely on the pilot's ability to monitor altitude and speed trends, ensuring oscillations settle rapidly under manual control. Autopilot systems and laws provide automated phugoid suppression through mechanisms that exceed natural capabilities. augmentation systems (SAS) incorporate phugoid dampers, which command deflections proportional to vertical velocity or , effectively increasing the mode's ratio to near-critical levels. In full authority digital engine (FADEC) integrated autopilots, -attitude-hold modes use proportional gains on displacement (Kθ) and (Kq) to overdamp the phugoid, transforming it into non-oscillatory and plunge responses; for instance, Kθ values around 20 can result in -mode frequencies approaching 5 Hz while maintaining . Automatic flight (AFC) modes, including altitude-hold with gains like Kh ≈ 5 × 10^{-5}, further minimize perturbations by coupling speed and altitude , often rendering phugoid imperceptible. modulation via auto-throttle servos, responding to velocity and , adds an additional layer by stabilizing states. Modern advancements in unmanned aerial vehicles (UAVs) and electric vertical takeoff and landing () aircraft leverage active control for robust phugoid suppression, often integrating Kalman filters for precise state estimation amid noisy sensor data. These filters fuse inertial, GPS, and air data to estimate and , enabling model predictive controllers or loops to preemptively adjust control surfaces and distributed propulsion, achieving damping ratios well above 0.5 even in gusty conditions. In designs, such systems ensure mode suppression during transition flights, where variable rotor configurations could otherwise exacerbate low-frequency oscillations. Testing protocols validate phugoid mitigation through iterative experiments to derive stability derivatives (e.g., lift curve slope and coefficients) and high-fidelity flight simulators to assess mode responses under certification maneuvers. Compliance with (FAR) Part 25 §25.181 requires that the phugoid motion not increase in amplitude during the first half of its natural period, regardless of initial amplitude, and not exhibit any dangerous characteristics between 1.13 V_{SR} and maximum speed. For handling qualities, military standards such as MIL-STD-1797B specify a minimum ratio of 0.04 for Level 1 flying qualities to avoid excessive pilot workload. These evaluations confirm margins across the , ensuring interventions like remain effective post-design.

Historical and Practical Context

Discovery and Early Research

Early observations of what would later be identified as phugoid motion emerged during 19th-century glider experiments. German aviation pioneer , conducting over 2,000 flights in his and gliders between 1891 and 1896, reaching distances up to 350 meters, highlighted the need for inherent stability in unpowered flight through observations of flight path variations. The further evidenced similar phenomena in their powered flights starting in 1903. During initial tests of the at , and Wilbur noted recurring speed-altitude trades where disturbances in pitch led to prolonged oscillations in forward speed and height, complicating control without active pilot intervention. Their diaries and correspondence described these as "porpoising" effects, influenced by the configuration and variable wind conditions, prompting iterative designs for better longitudinal . Formal identification of the phugoid as a distinct longitudinal mode occurred in the early through theoretical work. British aerodynamicist Frederick W. Lanchester coined the term "phugoid" in his 1908 Aerodonetics, deriving it from phygē (flight or fleeing) and the -oid (form or resemblance), to describe the long-period involving energy exchange between kinetic and potential forms. Lanchester's model, based on ballistic trajectories and glider analogies, approximated the motion as a simple harmonic wave independent of specifics, providing the first mathematical framework for its period scaling with flight speed. In 1911, mathematician G. H. Bryan formalized the mathematical description of the phugoid alongside the short-period mode, establishing the foundations for linearized stability analysis. Key theoretical advancements followed in the . By the 1940s, NACA engineer William H. extended these ideas through experimental studies on flying qualities, demonstrating in wind-tunnel tests how phugoid varied with center-of-gravity position and control surface effectiveness. ' 1948 report emphasized predictive criteria for mode separation, linking undamped phugoids to pilot workload in . Milestones in understanding phugoid motion included systematic NACA investigations in the . Reports like NACA-TR-96 (1921) on statical longitudinal stability quantified trim sensitivities to pitch disturbances, revealing oscillatory tendencies in power-off flight through free-flight model tests. These evolved into dynamic analyses by the decade's end, identifying phugoid-like responses in full-scale gliders. Post-World War II, digital simulations at (then NACA) in the enabled precise mode separation; early analog computers modeled nonlinear couplings, showing phugoid periods of 20-60 seconds in jet prototypes and confirming Lanchester's approximations under powered conditions. For instance, 1950s Langley studies used electronic differential analyzers to simulate F-86 responses, isolating phugoid from short-period modes for stability augmentation design. Early research faced significant gaps, particularly in distinguishing phugoid from short-period modes. Pre-1930s analyses often conflated the two due to coupled pitch-rate and altitude effects, leading to overestimations of overall in manual computations. Without computational tools until the , theoretical models relied on linear approximations that masked phugoid's weak , resulting in surprises during flight tests where long-period divergences appeared after initial short-period settling.

Notable Aviation Incidents

One notable incident involving phugoid motion occurred on April 4, 1975, during , when a U.S. Lockheed C-5A Galaxy (serial 68-0218) experienced a structural shortly after takeoff from in Saigon, . The of the aft cargo compartment bulkhead due to explosive decompression severed hydraulic lines to the tail, leaving pilots with limited pitch via engine throttles and roll via ailerons. This triggered severe phugoid oscillations, with cycles lasting approximately 60 seconds, as the aircraft traded altitude and speed uncontrollably during the attempted . Contributing factors included the inability to damp the motion effectively and crew fatigue from the high-workload scenario. The aircraft crash-landed in a rice paddy, resulting in 138 fatalities out of 314 people on board, primarily orphans being evacuated. The U.S. investigation highlighted the phugoid's role in the loss of and led to modifications, including reinforced cargo doors and improved hydraulic redundancy to prevent similar . Another significant case was Flight 123 on August 12, 1985, a 747SR-100 that suffered a rear bulkhead rupture 12 minutes after takeoff from Tokyo Haneda Airport, destroying the and all four hydraulic systems. The resulting loss of control surfaces excited initial phugoid oscillations, which amplified into undamped cycles lasting up to 90 seconds each, combined with , persisting for about 32 minutes as the crew struggled with asymmetric thrust. Triggers included an improper repair of the bulkhead from a prior incident, while contributing factors were the complete absence of hydraulic and pilot efforts to manage the motion using engine power alone. The aircraft crashed into , killing 520 of 524 aboard in the deadliest single-aircraft accident in history. The Japanese Aircraft Accident Investigation Commission's report, summarized by the FAA, emphasized phugoid dynamics in the prolonged loss-of-control sequence and prompted global regulatory changes, such as stricter maintenance protocols for fuselage structures and enhanced phugoid requirements in design standards. Phugoid motion also featured prominently in the July 19, 1989, crash of , a McDonnell Douglas DC-10-10 that lost all following an uncontained failure of its No. 2 engine fan disk over . This rendered primary flight controls inoperative, initiating slow vertical phugoid cycles characteristic of total hydraulic loss, with the aircraft oscillating in pitch while yawing rightward. The trigger was the engine explosion severing hydraulic lines, amplified by the lack of damping, and exacerbated by crew challenges in using differential engine thrust for attitude control. Despite a heroic crash landing at Sioux Gateway Airport, 112 of 296 people died. The NTSB investigation underscored phugoid oscillations as a key element in the loss-of-control progression and recommended advancements in hydraulic system independence, leading to industry-wide adoption of three independent hydraulic circuits in large transports and simulator training modules focused on phugoid recognition and engine-based recovery techniques. These incidents, investigated by bodies like the NTSB and international commissions, illustrate phugoid's contribution to loss-of-control events, though such cases remain relatively rare, comprising a small fraction of longitudinal stability-related accidents due to built-in . Post-accident reforms included FAA-mandated enhancements to phugoid criteria in standards (e.g., FAR Part 25), ensuring oscillations within specified limits, and integrated simulator curricula for pilots to identify and mitigate undamped phugoid growth from or control disruptions.

References

  1. [1]
    [PDF] Flight Stability and Automatic Control - Iowa State University
    Lanchester called the oscillatory motion the phugoid motion. He wanted to use the Greek word meaning "to fly" to describe his newly discovered motion; actually,.
  2. [2]
    [PDF] Notes on the Control of an Aircraft with Throttle Only
    Note that the application of throttle. causes a lightlydamped oscillation, called the phugoid mode. In this case, the period of the phugoid mode is about 1 ...
  3. [3]
    [PDF] Dynamic Stability
    The other mode has a much longer period and is rather lightly damped; this is called the phugoid mode. We illustrate this response using the stability ...
  4. [4]
  5. [5]
  6. [6]
    Find your center - AOPA
    Apr 1, 2017 · Phugoid oscillations typically have long periods of 20 to 100 seconds. These are pitch oscillations at almost constant angle of attack with the ...
  7. [7]
    [PDF] Phugoid characteristics of a YF-12 airplane with variable-geometry ...
    Each phugoid maneuver had between two and three cycles of free oscillation. Roll attitude was controlled by the roll autopilot to preclude inadvertent inputs to ...
  8. [8]
    [PDF] Simulation and Analysis of Aircraft Response Characteristics
    For longitudinal dynamic stability, the aircraft is said to have two dynamic modes which are called short period mode and phugoid mode. These are different ...
  9. [9]
    Aircraft Modes of Motion
    The Phugoid is an oscillation with almost no angle of attack change. There is little damping, and this mode can be unstable but the aircraft response is usually ...
  10. [10]
    Ask Us - Lift, Wind & Porpoising - Aerospaceweb.org
    Jun 19, 2005 · A phugoid is often created purposefully when a pilot rapidly pitches a plane's nose up or down and then takes his hands off the controls. The ...
  11. [11]
    [PDF] Lecture AC-1 Aircraft Dynamics
    motion of an aircraft (called the Phugoid mode) using a very simple balance ... • Predict the changes to the aerodynamic forces and moments using a first.
  12. [12]
    [PDF] Dynamics of Flight - Stability and Control - aerocastle
    Etkin, Bernard. Dynamics of flight : stability and control 1 Bernard Etkin, Lloyd. Duff Reid.-3rd ed. p. cm.
  13. [13]
    [PDF] Linearized Longitudinal Equations of Motion - Robert F. Stengel
    Nov 13, 2018 · Effects of phugoid perturbations on phugoid motion. Effects of phugoid perturbations on short-period motion. Effects of short-period.
  14. [14]
    [PDF] Lecture AC 2 Aircraft Longitudinal Dynamics
    • Longitudinal equations (1–15) can be rewritten as: m˙u = Xuu + Xww − mg ... – There is no roll/yaw motion, so q =˙θ. – The control commands ∆Xc ...
  15. [15]
    [PDF] Longitudinal Dynamics
    – Phugoid peaks present, but short period mode is very weak (not in u, low in γ, w). ⇒ entirely consistent with the step response. – Thrust controls speed ( ...
  16. [16]
    Phugoid Approximation for Conventional Airplanes | Journal of Aircraft
    Phugoid Approximation for Conventional Airplanes. W. F. Phillips, AIAA Journals, 2012. Phugoid Approximation Revisited. S. Kamesh, AIAA Journals, 2012. Unified ...
  17. [17]
    Aircraft Stability & Control – Introduction to Aerospace Flight Vehicles
    Positive static stability indicates that the aircraft returns to its original flight condition, while neutral static stability means it remains in the new ...
  18. [18]
    [PDF] A theoretical analysis of airplane longitudinal stability and control as ...
    On the other hand, it was found that positive wind shear can cause the phugoid to become aperiodic and unstable. In this case, a stability boundary for the ...
  19. [19]
    Dynamic Characteristic and Control of a Hypersonic Flight Vehicle
    The longitudinal dynamics of hypersonic flight vehicles present an unstable phugoid mode and a new height mode. Hypersonic flight will be subject to ...
  20. [20]
    [PDF] Conceptual Design Optimization of an Augmented Stability Aircraft ...
    Augmenting the aircraft's stability through active control will allow the stabilizers to be reduced in size thus reducing the total wetted area of the aircraft.
  21. [21]
    Phugoid Mode - an overview | ScienceDirect Topics
    Phugoid mode refers to a specific oscillatory motion of an aircraft in which it experiences periodic variations in altitude and airspeed, ...
  22. [22]
    Dynamic Stability - Aviation Safety Magazine
    Because of its long duration, the phugoid mode is easily controlled by the pilot. But a lightly damped phugoid will require more pilot inputs and thus can ...
  23. [23]
    [PDF] THE EFFECTS OF AN AUTOPILOT ON AIRPLANE RESPONSES TO ...
    Its motion is described by the longi- tudinal degrees of freedom: forward speed, plunge, and pitch, featuring phugoid and short- period modes. The automatic- ...
  24. [24]
    [PDF] Stability augmentation systems - Aerostudents
    Stability augmentation systems make the aircraft more stable. There are SASs for both the dynamic sta- bility (whether the eigenmotions don't diverge) and ...Missing: strategies | Show results with:strategies
  25. [25]
    [PDF] Phugoid Damping Control - DTIC
    Mar 20, 2001 · Phugoid damping control entails the design of compensators for the control of the aircraft's slow states. Thus, in the pitch channel, one ...
  26. [26]
    Thrust-Control System for Emergency Control of an Airplane
    Mar 28, 2017 · This Auto-Throttle Servocontrol System responds to velocity and pitch-angle feedback by modulating engine thrust to damp the phugoid oscillation and maintain a ...<|control11|><|separator|>
  27. [27]
    [PDF] Implementation Of Flight Control System Based On Kalman And PID ...
    The objectives of this research are to design UAV formation flight control system using combination of Kalman & PID, to study regulation quality variation under ...Missing: modern suppression eVTOL
  28. [28]
    [PDF] Advances in Aero-Propulsive Modeling for Fixed-Wing and eVTOL ...
    Apr 25, 2023 · This dissertation presents several mathematical modeling research advancements for these modern aircraft that foster accurate description and ...
  29. [29]
  30. [30]
    What is an acceptable phugoid oscillation duration?
    Nov 20, 2017 · MIL-STD-1797B specifies a damping ratio of at least 0.04 for Level 1 handling quality. The typical phugoid period is around 90 seconds.Why does static directional stability decrease with altitude?Why are the Phugoid and Short period approximated instead of ...More results from aviation.stackexchange.com
  31. [31]
    [PDF] Flight Controls of Otto Lilienthal's Experimental Monoplane from 1895
    Lilienthal stated that for landing, similar to turning, the glider requires a counterintuitive pilot motion (see Fig. 3, right-hand side). He reported ...Missing: undulatory phugoid
  32. [32]
    1903-The First Flight - Wright Brothers - National Park Service
    Oct 10, 2025 · The 27-mph wind was harder than they would have liked, since their predicted cruising speed was only 30-35 mph. The headwind would slow ...Missing: altitude trades phugoid<|separator|>
  33. [33]
    [PDF] F.W. Lanchester and the Great Divide.pdf - Royal Aeronautical Society
    Lanchester continued to work on his Phugoid Theory, and in the process came upon the experiments of Pénaud, Mouillard and Marey, all of whom had experimented ...
  34. [34]
    [PDF] Chapter 3 The Art of Phugoid
    Since the Greek root phug (φä7η, pronounced "fyoog") as well as the Latin root fug- both correspond to the English word "flight", he decided on the name " ...Missing: oid | Show results with:oid
  35. [35]
    [PDF] Theodore von Kármán
    His work on the buckling of elastic structures has become a standard part of the engineering theory of elastic stability. His work on plasticity and plastic ...
  36. [36]
    [PDF] Journey in Aeronautical Research
    My father,. William Phillips, was employed by this com- pany. His background was in chemistry and prior to working at Lever Brothers, he had been employed in ...
  37. [37]
    [PDF] Airplane Stability and Control, Second Edition - RexResearch1
    R. C. Nelson, Flight Stability and Automatic Control. McGraw-Hill, 1989 ... both the phugoid or long-period mode and the height mode. Thrust effects ...
  38. [38]
    A Study on Aircraft Longitudinal Stability | Aerotecnica Missili & Spazio
    Feb 15, 2022 · “Statical longitudinal stability of airplanes”, NACA-TR-96, National Advisory Committee for Aeronautics, 1921. Reports of National Advisory ...Missing: 1920s | Show results with:1920s
  39. [39]
  40. [40]
    [PDF] The birth of airplane stability theory - SciSpace
    [39] Phillips, W.F. Phugoid Approximation for Conventional Airplanes. Journal of Aircraft,. Vol. 37, Nº 1, 2000, pp. 30-36. [40] Pamadi, B.N. Performance ...<|control11|><|separator|>
  41. [41]
    [PDF] Touchdown: - NASA
    These strips also dominate certain unbalanced motions no pilot wants unleashed, motions known as the phugoid oscillation and the Dutch roll oscillation. You ...Missing: notable | Show results with:notable
  42. [42]
    [PDF] AIRCRAFT ACCIDENT INVESTIGATION REPORT Japan Air Lines ...
    This Report on Japan Air Lines JA8119 has been prepared by Aircraft. Accident Investigation Commission in accordance with Article of Aircraft. Accident ...