Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2025 Jul 2;15:23382. doi: 10.1038/s41598-025-07639-6

InvSim algorithm for pre-computing airplane flight controls in limited-range autonomous missions, and demonstration via double-roll maneuver of Mirage III fighters

Osama A Marzouk 1,
PMCID: PMC12222780  PMID: 40603517

Abstract

In this work, we start with a generic mathematical framework for the equations of motion (EOM) in flight mechanics with six degrees of freedom (6-DOF) for a general (not necessarily symmetric) fixed-wing aircraft. This mathematical framework incorporates (1) body axes (fixed in the airplane at its center of gravity), (2) inertial axes (fixed in the earth/ground at the take-off point), wind axes (aligned with the flight path/course), (3) spherical flight path angles (azimuth angle measured clockwise from the geographic north, and elevation angle measured above the horizon plane), and (4) spherical flight angles (angle of attack and sideslip angle). We then manipulate these equations of motion to derive a customized version suitable for inverse simulation flight mechanics, where a target flight trajectory is specified while a set of corresponding necessary flight controls to achieve that maneuver are predicted. We then present a numerical procedure for integrating the developed inverse simulation (InvSim) system in time; utilizing (1) symbolic mathematics, (2) explicit fourth-order Runge–Kutta (RK4) numerical integration technique, and (3) expressions based on the finite difference method (FDM); such that the four necessary control variables (engine thrust force, ailerons’ deflection angle, elevators’ deflection angle, and rudder’s deflection angle) are computed as discrete values over the entire maneuver time, and these calculated control values enable the airplane to achieve the desired flight trajectory, which is specified by three inertial Cartesian coordinates of the airplane, in addition to the Euler’s roll angle. We finally demonstrate the proposed numerical procedure of flight mechanics inverse simulation (InvSim) through an example case that is representative of the Mirage III family of French fighter airplanes, in which a straight subsonic flight with a double-roll maneuver over a duration of 30 s at an altitude of 5 km (3.107 mi or 16,404 ft) is inversely simulated.

Keywords: Fixed-wing aircraft, Airplane, Flight mechanics, Inverse simulation, Maneuver, Trajectory, RK4, FDM, Mirage III

Subject terms: Applied mathematics, Aerospace engineering

Introduction

Mathematical modeling and numerical simulations are important tools for describing various nonlinear complex phenomena and processes, as well as implementing computer-aided design (CAD), computational fluid dynamics (CFD), and automatic control110. Flight mechanics (also called flight dynamics) is one of the engineering fields that benefit largely from mathematical modeling and numerical simulation; because simple analytical reduced-order solutions in aerospace applications and aerodynamics are typically not available except under very restrained conditions; and experimental techniques through wind-tunnel tests (WTT) and flight tests are stochastic (non-deterministic), expensive, limited in terms of the amount of data that can be measured directly, and sometimes intrusive (influencing the domain being tested)1120.

Inverse simulation (InvSim) in flight mechanics is a normative category of flight mechanics modeling in which a desired flight trajectory (flight maneuver or flight mission) is specified through a number of inputs, while the corresponding flight controls (the model-based feedforward control variables) needed to achieve this trajectory are predicted; the opposite of this modeling process is called forward flight mechanics simulation, which is an exploratory category of flight mechanics2127. Accurate inverse simulation flight mechanics modeling facilitates advanced modes of transportation through pilotless (autonomous) aviation activities, such as regular electrified urban air mobility (e-UAM) trips within smart cities or between neighbor cities, powered by clean renewables; although additional real-time control systems for external perturbation suppression should be augmented2832.

This work is a sequel to a previous part in which we provided a detailed mathematical framework for general modeling of the motion of a fixed-wing aircraft, and this framework has several advantages; namely: (1) all six degrees of freedom (6-DOF) are included, (2) the singularity of upward/downward vertical flight (encountered in a traditional Euler-angle representation) is avoided, (3) linear (translational) momentum equations are transformed such that the order of magnitude of adjacent terms does not suffer from large discrepancy under fast dynamics, (4) the assumption of airplane symmetry is eliminated, (5) the aerodynamic details for all the flight-dependent aerodynamic/stability coefficients are clearly expressed, (6) three sets of axes (namely, inertial ground/earth axes, body-fixed axes, and wind axes) are utilized to describe the attitude of the airplane, rather than using the Euler angles solely, (7) two flight-path angles (azimuth and elevation) are utilized as an intermediate spherical coordinate system; allowing the separation of the flight path direction relative to air (this is the airplane “course”) from the airplane attitude relative to the ground (this is the airplane “heading”), (8) only scalar equations (rather than vector equations or quaternions) are utilized, which simplifies the implementation process as a computer-based flight mechanics simulator, (9) the variation of the air density with altitude is accounted for, using the international standard atmosphere (ISA) model for air as an ideal homogeneous gas, and (10) the model can be easily adjusted to specific airplane conditions through user-defined input parameters or specialized airplane features (such as a nonlinear lift coefficient profile, instead of linear dependence on the angle of attack) can be handled through minor modifications3356.

In the current work, the previously explained general nonlinear differential–algebraic equations (DAE) system for flight mechanics is further reformulated such that they fit specifically inverse simulation flight mechanics, and a detailed computational algorithm is presented to numerically integrate these equations of motion (EOM) through explicit mathematical expressions that do not require solving an algebraic system of equations, with the use of the fourth-order Runge–Kutta (RK4) numerical integration method; also numerical differentiation expressions based on the finite difference method (FDM) may be used5764.

We implement this proposed numerical procedure as a computer code, and demonstrate its utilization while inversely simulating a continuous-double-roll maneuver with a set of airplane data that nearly corresponds to the Mirage III fighter aircraft, produced by the French aerospace company “Dassault Aviation”65,66.

Research method

Problem statement

The inverse simulation (InvSim) flight mechanics algorithm proposed here considers the airplane as a MIMO (multi-input multi-output) control system, with four inputs and four outputs. The four inputs are specified as the inertial (ground-referenced) Cartesian coordinates (Inline graphic) of the airplane’s center of gravity (CG), and these coordinates may be described as analytical functions of time or as discrete values recorded at specified time values, uniformly spaced with a constant time step (Inline graphic. The (Inline graphic) coordinate represents the signed displacement traveled by the airplane in the geographic north direction from the initial flight point (the take-off point). The (Inline graphic) coordinate represents the signed displacement traveled by the airplane in the geographic east direction from the initial flight point. The (Inline graphic) coordinate represents the signed displacement traveled by the airplane toward the earth’s center from the initial flight point, and thus this coordinate is expected to have negative values except during parts of the maneuver where the airplane descends to an altitude below the initial altitude. The three inertial (ground-referenced) Cartesian coordinates (Inline graphic) are expressed in meters (m). Figure 1 illustrates these three input coordinates. The term “gravity axis” here refers to the inertial axis pointing toward the earth’s center (perpendicular to the horizon plane), and it is opposite to the “altitude axis” that is also perpendicular to the horizon plane but points toward the sky. The origin of these ground-referenced rectangular coordinates is the initial flight point (the first location of the trajectory to be inversely simulated).

Fig. 1.

Fig. 1

Illustration of the three inertial (ground-referenced) coordinates, which are three of the four inputs to InvSim.

In the current work, the altitude of the airplane’s center of gravity as measured from the mean sea level (MSL) is designated by the symbol (Inline graphic), and the initial altitude is designated by the symbol (Inline graphic). Therefore, the gained height of the airplane’s center of gravity above the initial trajectory point is (Inline graphic); which should be equal to the negative value of the (Inline graphic) coordinate. Therefore,

graphic file with name 41598_2025_7639_Article_Equ1.gif 1

The fourth input to the proposed InvSim model is the roll angle (expressed in radians, “rad”), which is one of the three Euler angles, and it describes the lateral attitude of the airplane. If the starboard (right) tip and the port (left) tip of the wing are at the same vertical position (having the same altitude), then the roll angle (Inline graphic) is zero. By convention, the roll angle is positive if the wing’s starboard/right tip tilts down and the wing’s port/left tip tilts up67,68. Figure 2 illustrates the roll angle in a particular configuration where it has a positive value.

Fig. 2.

Fig. 2

Illustration of the roll angle, which is the fourth input to InvSim.

The four outputs from the proposed inverse simulation (InvSim) flight mechanics model here are the four flight controls (four control variables), computed as four discrete series of values (four numerical vectors). These controls are in charge of adjusting the speed and orientation of the airplane, such that the input conditions are satisfied. These four output controls are:

  1. The engine thrust force (Inline graphic), in newtons (N). This control can also be an electric propulsion force in the case of using an electric propeller or an electric ducted fan (EDF) instead of a traditional fuel-fired heat engine6974, and this electric propulsion7581 (or the use of heat engines powered by clean alternative non-fossil fuels8285) has an environmental advantage due to eliminating greenhouse (GHG) emissions as a results of burning a fossil fuel8689. While the presented InvSim model theoretically can admit thrust reversal (TR) or reverse thrust (corresponding to negative thrust values) during brief periods of the whole trajectory, this is typically not acceptable in fixed-wind airplanes9093.

  2. The ailerons’ deflection angle (Inline graphic), in radians (rad). This deflection is primarily in charge of the roll degree of freedom. This deflection angle is positive when the hinged starboard/right aileron tilts up and simultaneously the hinged port/left aileron tilts down (which induces a rolling moment leading to a positive roll angle, Inline graphic).

  3. The elevators’ deflection angle (Inline graphic), in radians (rad). This deflection is primarily in charge of the pitch degree of freedom. This deflection angle is positive when both hinged elevators tilt down (which induces a pitching moment leading to a positive pitch angle, Inline graphic, where the airplane’s nose tilts up).

  4. The rudder’s deflection angle (Inline graphic), in radians (rad). This deflection is primarily in charge of the yaw degree of freedom. This deflection angle is positive when the hinged rudder tilts toward the starboard/right side (which induces a yawing moment leading to a positive yaw angle “heading angle”, Inline graphic, where the airplane’s nose tilts toward the starboard/right side).

The sign convention of the three control deflection angles (Inline graphic) are positive when the resulting body-referenced rotations become related to their respective body axes (the longitudinal or forward axis, the lateral or starboard axis, and the bottom or third axis) through the corkscrew rule (also called the curl right-hand rule or the right-hand grip rule)94,95.

Figure 3 illustrates these four flight controls (which are the four main outputs from the InvSim model).

Fig. 3.

Fig. 3

Illustration of the four flight controls.

Research approach

In order to solve the stated problem in the previous subsection, symbolic mathematical manipulation is combined with computational methods to build and present here a proposed numerical algorithm for solving the inverse simulation flight mechanics problem. A general set of flight mechanics equations of motion are first presented. These equations are then reformulated to be in the inverse simulation InvSim form, such that the four output flight controls can be obtained for specified inputs (trajectory coordinates and roll angle). Then, a proposed algorithm is presented, which allows numerically integrating the reformulated coupled differential–algebraic equations (DAE) in a simple way that can be implemented using a general-purpose computer programming language, without the need for specialized capabilities or commercial packages96103. We here use the MATLAB/Octave programming language, which is a high-level interpreted language suitable especially for numerical computations, although other programming languages (such as Python or FORTRAN) can also be used104111.

The success of the proposed computational InvSim algorithm and the underlying mathematical model is supported by numerically simulating an example flight maneuver. We obtained some symbolic expressions for needed derivative terms manually using normal calculus rules; and independently using Mathematica (a popular software tool for symbolic mathematics and computations that has been used in many research works before), and both sets of obtained expressions were exactly compatible112122.

General equations of motion

In this section, we present part of a flight mechanics mathematical framework, which is not optimized specifically for inverse simulations. The equations of motion listed in this section are not just the six main equations (the three linear/translational momentum equations and the three angular/rotational momentum equations) for a six-degree-of-freedom rigid body in flight. Instead, auxiliary intermediate equations are also listed, since they need to be considered along with the six main equations as a complete integrated coupled system.

Angular velocity vector in body axes

The roll rate (Inline graphic), pitch rate (Inline graphic), and yaw rate (Inline graphic) are related to the Euler angular rates or Euler rates (Inline graphic) that are the time derivatives of the Euler angles, according to

graphic file with name 41598_2025_7639_Article_Equ2.gif 2
graphic file with name 41598_2025_7639_Article_Equ3.gif 3
graphic file with name 41598_2025_7639_Article_Equ4.gif 4

Figure 4 illustrates the three Euler angles (Inline graphic) which describe the three-stage rotational transformation from the inertial axes system (global/earth north, global/earth east, and global/earth gravity) into the body-fixed axes system (longitudinal, lateral, bottom; or Inline graphic)123126.

Fig. 4.

Fig. 4

Illustration of the three Euler angles, as well as the inertial axes and the body axes.

Linear-momentum equations and equilibrium

The linear-momentum equations are transformed from the Cartesian system into the translating spherical wind axes, whose origin is translating with the body (coincides with the airplane’s center of gravity), and whose coordinates are the velocity magnitude (Inline graphic), the sideslip angle (Inline graphic), and the angle of attack (AoA or Inline graphic). These three variables (Inline graphic) can be viewed as spherical coordinates, with the velocity magnitude (Inline graphic) being the radial coordinate. The forward wind axis Inline graphic is tangent to the flight path and thus coincides with the total velocity vector (this is the velocity of the airplane’s center of gravity). The sideslip angle (Inline graphic) represents a rotational transformation of the wind axes such that the rotated Inline graphic lies in the airplane’s fixed plane Inline graphic (the airplane’s midplane), which is the plane of symmetry for symmetric airplanes.

The condition (Inline graphic = 0) means that the incoming airflow relative to the airplane is symmetric with respect to the airplane’s midplane. The angle of attack (Inline graphic) represents a subsequent rotational transformation (after the Inline graphic rotational transformation) such that the twice-rotated wind axes (Inline graphic) coincide with the body-fixed axes (Inline graphic). The components of the velocity vector (Inline graphic) along the body-fixed axes (Inline graphic) are (Inline graphic), respectively. Figure 5 illustrates the wind angles (Inline graphic and Inline graphic), as well as the relation between the wind axes system (Inline graphic) and the body axes system (Inline graphic).

Fig. 5.

Fig. 5

Illustration of the wind axes and their angles (sideslip angle and angle of attack).

Although our definition for the angle of attack (AoA or Inline graphic) discussed above, as an angle of axes transformation, implies that in the absence of any sideslip (Inline graphic = 0) and when Inline graphic = 0, the longitudinal body axis (Inline graphic) coincides with the forward wind axis (Inline graphic). This configuration corresponds to a horizontal steady-level flight with zero tilts by the airplane (the body-fixed plane Inline graphic is parallel to the horizon plane), and we refer to this configuration here as the “equilibrium” condition of flight. However, despite the horizontal orientation of the airplane’s body, and the zero value assigned to the angle of attack (Inline graphic) as per our definition, aerodynamic principles demand that there must be an exerted lifting force to counteract the weight of the airplane during an equilibrium flight (as well as an exerted thrust force to counteract the drag force resisting the forward motion)127129.

The equilibrium lift force may require either (1) using a nonsymmetrical (cambered) wing airfoil section while keeping the wing mounted parallel to the main airplane body (the fuselage), such that a non-zero upward lifting force can be exerted even when the airplane is flying horizontally through the ambient air with its nose-to-tail longitudinal axis is also oriented horizontal; or (2) mounting the wing (which can be cambered or symmetric in this case, not necessarily cambered) at a small tilt angle called the angle of incidence or the mounting angle (Inline graphic), such that when the airplane flies horizontally with a horizontal longitudinal body axis (a horizontal fuselage), there is still a non-zero lifting force that achieves the equilibrium condition because the tilted wing now faces the air asymmetrically, causing the air pressure at the lower surface of the wing to be higher than the air pressure at the upper surface of the wing due to the asymmetric air flow around the wing, and this leads to the upward lifting force component (combined with an aerodynamic drag force component)130143. In the current study, we adopt the latter choice; thus, we assume that the wing is installed into the fuselage at a small angle of incidence (Inline graphic) that satisfies the equilibrium condition; and this tilt angle (Inline graphic) is excluded from the angle of attack (Inline graphic) used in the modeling and numerical simulation discussed here.

In the flight mechanics model presented here, the angle of attack (Inline graphic) excludes any equilibrium angle between the velocity vector of the airplane’s center of gravity (whose magnitude is Inline graphic) and the wing’s mean chord line (the virtual straight line connecting the leading edge with the trailing edge of the wing of an airfoil section)144149. Therefore, the symbol (Inline graphic) that appears in this study means the change in the conventional angle of attack (denoted here by the symbol Inline graphic) from its equilibrium conventional value (Inline graphic), which is assumed to be equal to the angle of incidence in the current work. Therefore, we have

graphic file with name 41598_2025_7639_Article_Equ5.gif 5
graphic file with name 41598_2025_7639_Article_Equ6.gif 6
graphic file with name 41598_2025_7639_Article_Equ7.gif 7

In aeronautics, the equilibrium conventional angle of attack (Inline graphic) is not truly a constant. Rather, this equilibrium conventional angle of attack depends on the flight speed, and it also depends on the air density, which in turn depends on the air pressure150152 (as for any ideal compressible gaseous medium such as ambient air153157). This ambient air pressure drops with the flight altitude158160. Therefore, our assumption of equality between the equilibrium conventional angle of attack and the angle of incidence (Inline graphic) implies that the angle of incidence is effectively adjustable; for example through using movable flaps attached to the wing, and these allows adjusting the wing’s effective angle of incidence161163.

Figure 6 illustrates the difference between the angle of incidence (Inline graphic) and the angle of attack (Inline graphic). The figure also shows how they differ from the Euler pitch angle (Inline graphic) discussed in the previous subsection, and how they differ from the elevation flight path angle (Inline graphic) that describes the climb angle of the airplane based on its course of flight as a point particle.

Fig. 6.

Fig. 6

Illustration of four tilt angles in the midplane of the airplane.

The linear-momentum equations (with their components corresponding to the wind axes) are

graphic file with name 41598_2025_7639_Article_Equ8.gif 8
graphic file with name 41598_2025_7639_Article_Equ9.gif 9
graphic file with name 41598_2025_7639_Article_Equ10.gif 10

where Eq. (8) is the Inline graphic component of the vector linear momentum equation, Eq. (9) is its Inline graphic component, and Eq. (10) is its Inline graphic component.

In the above equations, (Inline graphic) is the airplane mass, (Inline graphic) is its wing planform area, (Inline graphic) is the thrust force, (Inline graphic) is the Euler’s pitch angle, (Inline graphic) is the gravitational acceleration, (Inline graphic) are nondimensional force coefficients (to be discussed later), and (Inline graphic) is the dynamic pressure defined as164166

graphic file with name 41598_2025_7639_Article_Equ11.gif 11

where (Inline graphic) is the air density.

The gravitational acceleration (Inline graphic) is treated here as a constant with the value of 9.81 m/s2. Therefore,

graphic file with name 41598_2025_7639_Article_Equ12.gif 12

It may be useful to add here that if the linear/translational equations of motion are formulated along the body-fixed axes; then their three components along the longitudinal axis (Inline graphic), the lateral body-fixed axis (Inline graphic), and the third/bottom body-fixed axis (Inline graphic); respectively; become

graphic file with name 41598_2025_7639_Article_Equ13.gif 13
graphic file with name 41598_2025_7639_Article_Equ14.gif 14
graphic file with name 41598_2025_7639_Article_Equ15.gif 15

These are the three components of the following vector linear-momentum equation in the body-fixed axes:

graphic file with name 41598_2025_7639_Article_Equ16.gif 16

where (Inline graphic) are the three components of the velocity vector along the three body-fixed axes (Inline graphic), respectively. The right-hand side of the above vector equation is the total applied forces on the airplane. The first vector term on the right-hand side represents the aerodynamic forces, the second vector term on the right-hand side represents the weight force, and the third vector term on the right-hand side represents the thrust force.

Angular-momentum equations

A derived geometric constant (Inline graphic) needs to be computed once, and this constant is the determinant of the symmetric inertia tensor. The three diagonal components of this tensor are the moments of inertia (or rectangular moments of inertia) for rotations perpendicular to the three corresponding body axes (centered at the airplane’s center of gravity), which are positive numbers for a rigid body; while the off-diagonal elements are the products of inertia, which can be negative, zero, or positive numbers167173.

The constant (Inline graphic) is expressed mathematically as

graphic file with name 41598_2025_7639_Article_Equ17.gif 17

The SI unit of each component of the inertia tensor is kg.m2, and thus the SI unit of (Inline graphic) is kg3.m6. The six components of the inertia tensor are further explained in Table 1.

Table 1.

List of the six independent components in the airplane’s symmetric inertia tensor.

Inertia symbol Alternative symbol(s) Meaning
A Inline graphic Body-referenced moment of inertia about the longitudinal axis (Inline graphic)
B Inline graphic Body-referenced moment of inertia about the lateral axis (Inline graphic)
C Inline graphic Body-referenced moment of inertia about the bottom/third axis (Inline graphic)
D Inline graphic Body-referenced product of inertia in the plane (Inline graphic)
E Inline graphic Body-referenced product of inertia in the plane (Inline graphic)
F Inline graphic Body-referenced product of inertia in the plane (Inline graphic)

For a symmetric airplane, the left (port) half is identical (but reflected) to the right (starboard) half; and in such a case of left–right symmetry, only the product of inertia in the symmetry midplane plane (Inline graphic) is non-zero (Inline graphic ≠ 0)174176.

Because the rotational equations of motion are expressed in the body-fixed axes and the above inertia terms are formulated also with respect to the body-fixed axes, these inertia terms are considered invariant constants in the rotational equations of motion.

Three auxiliary moments (Inline graphic) are defined through three algebraic equations as

graphic file with name 41598_2025_7639_Article_Equ18.gif 18
graphic file with name 41598_2025_7639_Article_Equ19.gif 19
graphic file with name 41598_2025_7639_Article_Equ20.gif 20

Finally, the main rotational equations of motion for the airplane about its body axes are

graphic file with name 41598_2025_7639_Article_Equ21.gif 21
graphic file with name 41598_2025_7639_Article_Equ22.gif 22
graphic file with name 41598_2025_7639_Article_Equ23.gif 23

In the special case of a symmetric airplane, the seven equations presented in the current subsection, namely Eqs. (1723), can be replaced by only three differential angular-momentum equations that are listed below; which can be derived from Eqs. (17–23) after setting (Inline graphic = 0) and (Inline graphic = 0) and performing some mathematical manipulation177185.

graphic file with name 41598_2025_7639_Article_Equ24.gif 24
graphic file with name 41598_2025_7639_Article_Equ25.gif 25
graphic file with name 41598_2025_7639_Article_Equ26.gif 26

These special simpler equations can be further manipulated and expressed using the alternative symbols for the inertia terms as

graphic file with name 41598_2025_7639_Article_Equ27.gif 27
graphic file with name 41598_2025_7639_Article_Equ28.gif 28
graphic file with name 41598_2025_7639_Article_Equ29.gif 29

Inertial velocity

The equations relating the rates of the ground-referenced inertial coordinates (Inline graphic) to the velocity magnitude and the spherical flight path angles (azimuth flight path angle Inline graphic, and elevation flight path angle Inline graphic) are

graphic file with name 41598_2025_7639_Article_Equ30.gif 30
graphic file with name 41598_2025_7639_Article_Equ31.gif 31
graphic file with name 41598_2025_7639_Article_Equ32.gif 32

Figure 7 illustrates the two spherical flight path angles (Inline graphic and Inline graphic), which describe the flying course (direction) of the airplane, treated as a particle, with respect to the inertial origin (the initial trajectory point) using the three spherical coordinates (Inline graphic) as an alternative to the Cartesian inertial velocity components (Inline graphic).

Fig. 7.

Fig. 7

Illustration of the two flight path angles.

Flight path angles

Two additional algebraic equations relate the two flight path angles (Inline graphic) to the three Euler angles (Inline graphic) and the two wind axes angles (Inline graphic, Inline graphic) are provided below186191.

graphic file with name 41598_2025_7639_Article_Equ33.gif 33
graphic file with name 41598_2025_7639_Article_Equ34.gif 34

From the two above equations; it can be proven that in the condition of equilibrium flight (Inline graphic = 0, Inline graphic = 0); the Euler yaw angle (Inline graphic), also called “heading angle”, becomes equal to the azimuth flight path angle (Inline graphic); and the Euler pitch angle (Inline graphic) becomes equal to the elevation flight path angle (Inline graphic).

Three aerodynamic forces

The three body-axes aerodynamic forces acting on the airplane are expressed as

graphic file with name 41598_2025_7639_Article_Equ35.gif 35
graphic file with name 41598_2025_7639_Article_Equ36.gif 36
graphic file with name 41598_2025_7639_Article_Equ37.gif 37

In the above equations, (Inline graphic) is the aerodynamic force along the longitudinal body-fixed axis, and its unit vector exactly coincides with the unit vector of the thrust vector; (Inline graphic) is the aerodynamic force along the starboard/right lateral body-fixed axis; and (Inline graphic) is the aerodynamic force along the bottom/third body-fixed axis.

Three moments

The total moment vector is resolved into three components (which we also refer to as “three moments”) along the body axes. These moments are expressed in terms of nondimensional moment coefficients (Inline graphic) as192194

graphic file with name 41598_2025_7639_Article_Equ38.gif 38
graphic file with name 41598_2025_7639_Article_Equ39.gif 39
graphic file with name 41598_2025_7639_Article_Equ40.gif 40

In the above equations, (Inline graphic) is the rolling moment about the longitudinal body-fixed axis (Inline graphic), (Inline graphic) is the pitching moment about the lateral body-fixed axis (Inline graphic), and (Inline graphic) is the yawing moment about the third/bottom body-fixed axis (Inline graphic). In addition, (Inline graphic), which is the mean aerodynamic chord (MAC) taken as a characteristic length for nondimensionalizing the pitching moment (longitudinal stability)195198.

The mean aerodynamic chord (MAC)199,200 is defined as follows:

graphic file with name 41598_2025_7639_Article_Equ41.gif 41

where (Inline graphic) is the local chord distance as a function of the lateral distance (Inline graphic), and (Inline graphic) is the semi-span. The mean aerodynamic chord (MAC) is different from the mean geometric chord (MGC), also called standard mean chord (SMC)201,202, which is defined as

graphic file with name 41598_2025_7639_Article_Equ42.gif 42

For most airplanes (with a wing planform resembling a rectangle or a trapezoidal “trapezium”), the values and spanwise location of the standard mean chord (SMC) and the mean aerodynamic chord (MAC) are close to each other; thus, they may be practically treated interchangeably203. However, for delta wings, it can be shown that

graphic file with name 41598_2025_7639_Article_Equ43.gif 43

where (Inline graphic) is the maximum local chord (at the wing root), while

graphic file with name 41598_2025_7639_Article_Equ44.gif 44

or

graphic file with name 41598_2025_7639_Article_Equ45.gif 45

For a rectangular wing, the local chord is uniform. Thus, the standard mean chord (SMC) and the mean aerodynamic chord (MAC) are exactly equal, or

graphic file with name 41598_2025_7639_Article_Equ46.gif 46

It is worth mentioning here that the wing has a nondimensional characteristic attribute, which is the aspect ratio (AR), defined as the square of the span divided by the wing area, or

graphic file with name 41598_2025_7639_Article_Equ47.gif 47

For a rectangular wing, the aspect ratio reduces to the span-to-chord ratio, or

graphic file with name 41598_2025_7639_Article_Equ48.gif 48

The other characteristic length (Inline graphic) pertains to the rolling moment (lateral stability) and the yawing moment (directional stability), and this characteristic nondimensionalization length is the wing span (wingspan)204208.

Aerodynamic and stability coefficients

The aerodynamic lift coefficient (Inline graphic), aerodynamic drag coefficient (Inline graphic), and aerodynamic side-force coefficient (Inline graphic) are wind-axes nondimensional quantities from which the body-axes aerodynamic coefficients (Inline graphic) can be obtained209216.

The lift coefficient is modeled here as being directly related to the conventional angle of attack (Inline graphic) as

graphic file with name 41598_2025_7639_Article_Equ49.gif 49

where (Inline graphic) is the lift coefficient at zero conventional angle of attack, and (Inline graphic) is the gain in the lift coefficient per unit increase in the angle of attack (when expressed in radians), and both values are treated as constant parameters.

As mentioned earlier in “Linear-momentum equations and equilibrium” section, the term “angle of attack” (Inline graphic) used in the proposed flight mechanics simulation modeling here is the change (either positive or negative) in the conventional angle of attack from its equilibrium value (Inline graphic), which is assumed to be equal to the angle of incidence for the wing (Inline graphic). Therefore, the lift coefficient (Inline graphic) is related to the modeling/simulation angle of attack (Inline graphic) as

graphic file with name 41598_2025_7639_Article_Equ50.gif 50

Regardless of the lift coefficient being expressed as a function of the conventional angle of attack (Inline graphic) as in Eq. (49), or being expressed as a function of the modeling/simulation angle of attack (Inline graphic) as in Eq. (50), it has the same slope (Inline graphic). At any given value of (Inline graphic), the angular difference (Inline graphic) is equal to the equilibrium value of the conventional angle of attack (Inline graphic; and it is computed from the force balance between the weight of the airplane (Inline graphic) and the equilibrium lift force (Inline graphic), where (Inline graphic) is the lift coefficient at equilibrium flight. Therefore,

graphic file with name 41598_2025_7639_Article_Equ51.gif 51

Combining the above defining equation for the equilibrium lift coefficient with Eq. (49) that relates the lift coefficient (Inline graphic) linearly with the conventional angle of attack (Inline graphic) gives

graphic file with name 41598_2025_7639_Article_Equ52.gif 52

This leads to an explicit expression for the conventional equilibrium angle of attack (Inline graphic as

graphic file with name 41598_2025_7639_Article_Equ53.gif 53

This conventional equilibrium angle of attack (Inline graphic is approximately constant if the dynamic pressure (Inline graphic) changes only slightly, and this implies that the air density and the flight speed are nearly unchanged, and this is a valid assumption in steady-level flight.

Figure 8 illustrates the linear dependence of the lift coefficient (Inline graphic) on the modeling/simulation angle of attack (Inline graphic), which is the default angle of attack in the current work. The figure also illustrates the linear dependence of the lift coefficient (Inline graphic) on the conventional angle of attack (Inline graphic)217220. This linear dependence is appropriate as long as the airplane is away from aerodynamic stall conditions221227.

Fig. 8.

Fig. 8

Illustration of the lift coefficient profile.

Through the drag polar relationship, the drag coefficient depends on the lift coefficient as228230

graphic file with name 41598_2025_7639_Article_Equ54.gif 54

where (Inline graphic) and (Inline graphic) are additional aerodynamic constants.

The side-force coefficient depends on the sideslip angle as

graphic file with name 41598_2025_7639_Article_Equ55.gif 55

After knowing the wind-axes aerodynamic coefficients (Inline graphic), the body-axes aerodynamic coefficients (Inline graphic) can be obtained using straightforward trigonometric projections as

graphic file with name 41598_2025_7639_Article_Equ56.gif 56
graphic file with name 41598_2025_7639_Article_Equ57.gif 57
graphic file with name 41598_2025_7639_Article_Equ58.gif 58

The body-axes total-moment coefficients (Inline graphic) are modeled as231

graphic file with name 41598_2025_7639_Article_Equ59.gif 59
graphic file with name 41598_2025_7639_Article_Equ60.gif 60
graphic file with name 41598_2025_7639_Article_Equ61.gif 61

where (Inline graphic) are nondimensional constant parameters, while (Inline graphic) are dimensional constant parameters having the unit of 1/rad, (Inline graphic) is the ailerons’ deflection angle in radians, (Inline graphic) is the elevators’ deflection angle in radians, and (Inline graphic) is the rudder’s deflection angle in radians.

Air density and speed of sound

The variation of the air density with the altitude (Inline graphic) is governed here by the International Standard Atmosphere (ISA) model. Up to an altitude of about 11,000 m above mean sea level, the troposphere layer of the air is present, in which the air temperature decreases linearly with altitude, while the density declines nonlinearly with altitude at a faster rate according to232239.

graphic file with name 41598_2025_7639_Article_Equ62.gif 62

where (Inline graphic) is the lapse rate magnitude, taken as 0.0065 K/m; and (Inline graphic) is the ideal gas constant for air, which is the universal molar gas constant divided by the molecular weight, and this gas constant for air is 287 J/kg.K240243. In the previous equation, the constant 1.225 is the standard sea level air density in kg/m3 (Inline graphic); and the constant 288.15 is the standard sea level absolute temperature in kelvins (Inline graphic), which corresponds to 15 °C244248. The standard sea level atmospheric pressure in the ISA model is 101,325 Pa (101.325 kPa, 1.01325 bar, or 760 mmHg)249252.

Although the ISA model is based on geopotential altitudes for estimating the air density, we here utilize the geometric (true or orthometric) altitude253265. This simplification is aligned with our treatment of the gravitational acceleration as a universal constant. We assessed the difference in the two types of altitudes, and we found that the difference in the troposphere layer of interest here is small, as shown in Table 2.

Table 2.

Examples of the geometric and geopotential altitudes within the troposphere layer.

Geometric altitude (m) Geopotential altitude (m) Absolute difference (m) Percentage deviation
5,000.000 4,996.079 3.921 0.0785%
5,003.927 5,000.000 3.927 0.0785%
10,000.000 9,984.328 15.672 0.1568%
10,015.721 10,000.000 15.721 0.1571%

The percentage deviations in the above table are very small, much less than 1%. These percentage deviations were computed as266

graphic file with name 41598_2025_7639_Article_Equ63.gif 63

The relationship between the geometric altitude (Inline graphic) and geopotential altitude (denoted by Inline graphic) is

graphic file with name 41598_2025_7639_Article_Equ64.gif 64
graphic file with name 41598_2025_7639_Article_Equ65.gif 65

where (Inline graphic) is the mean radius of the earth, and it is taken here as 6,371,000 m 267271.

In Table 3, we further assess the influence of using a geometric altitude rather than a geopotential altitude when using Eq. (62) for computing the air density. At a geometric altitude of 5,000 m, for example; the air density should be computed strictly speaking at the slightly-lower geopotential altitude of approximately 4,996 m. Similarly, the air density computed at a geopotential altitude of 5,000 m strictly speaking corresponds to the slightly higher geometric altitude of approximately 5,004 m. However, the values in the table show that these three densities are nearly the same, and the error incurred by the simplifying assumption of using the geometric altitude in lieu of the geopotential altitude leads to a marginal error in the air density that is below 0.05% at altitudes near 5,000 m; corresponding approximately to the middle of the troposphere layer of the atmosphere.

Table 3.

Quantification of the error in air density calculation at altitudes near 5,000 m.

Air density computed using an altitude of 4996 m (Inline graphic) 0.736191 kg/m3
Air density computed using an altitude of 5000 m (Inline graphic) 0.735872 kg/m3
Air density computed using an altitude of 5004 m (Inline graphic) 0.735553 kg/m3
Inline graphic 0.0433%
Inline graphic –0.0433%

Similarly, Table 4 assesses the expected error in the air density but at higher altitudes near 10,000 m; located close to the upper end of the troposphere layer of the atmosphere. Although the error grows (nonlinearly) as the altitude increases, it remains very small, below 0.2%. Regarding the three altitudes listed in this table (9,984 m; 10,000 m; and 10,016 m); the geopotential altitude corresponding to a geometric altitude of 10,000 m is approximately 10,016 m; while the geometric altitude corresponding to a geopotential altitude of 10,000 m is approximately 9,984 m.

Table 4.

Quantification of the error in air density calculation at altitudes near 10,000 m.

Air density computed using an altitude of 9984 m (Inline graphic) 0.413234 kg/m3
Air density computed using an altitude of 10,000 m (Inline graphic) 0.412415 kg/m3
Air density computed using an altitude of 10,016 m (Inline graphic) 0.411597 kg/m3
Inline graphic 0.1986%
Inline graphic –0.1983%

Many commercial transport airplanes fly at cruising altitudes below 11 km, thus the troposphere formula is adequate for them272280. For higher altitudes, as in some military aircraft, another expression for the air density should be used, which corresponds to the tropopause layer of the atmosphere; it is an interspheric layer lying between the lower troposphere layer and the upper stratosphere layer, extending approximately between the altitudes 11,000 m and 20,000 m281286. In the tropopause layer of the atmosphere, the temperature is treated as constant (with a value of –56.50 °C or 216.65 K), making the tropopause an isothermal layer. Therefore, the decline of the air density with the altitude within the tropopause layer follows a different profile than the one described earlier for the non-isothermal troposphere layer. This decline is described as

graphic file with name 41598_2025_7639_Article_Equ66.gif 66

where the value (11,000) is the altitude (in meters) at the bottom edge of the tropopause layer, which is also the upper edge of the troposphere layer; the value (0.3636309) is the air density at this bottom edge of the tropopause layer (Inline graphic) in kg/m3, as computed by the tropospheric density equation; and the value (216.65) is the absolute temperature (in kelvins) of air within the tropopause layer (Inline graphic).

It is worth mentioning that the decline of the air density in the non-isothermal troposphere layer is slower than its decline in the isothermal tropopause layer, because the drop of temperature in the non-isothermal troposphere layer has a compressive effect on air, causing its density to tend to increase; however, the drop in the pressure due the diminishing weight of the above air column causes a stronger expansive effect of decreasing the air density; and resultantly the air density declines with the altitude287291. In the isothermal tropopause layer, the compressive temperature effect is eliminated, thereby magnifying the expansive pressure effect, and thus the air density declines faster with the altitude. To demonstrate this faster rate of density decline in the isothermal tropopause layer, applying Eq. (66) describing the tropopause’s air density to an altitude of zero (Inline graphic= 0), which belongs to the troposphere layer – outside the tropopause layer, gives an extrapolated benchmarking value for the sea-level air density of 2.0624 kg/m3, which clearly exceeds the true sea level value of 1.225 kg/m3292294.

Although the speed of sound (Inline graphic) in air is not an essential variable in the presented flight mechanics model, it is still an important property in aeronautics, because the ratio between the airplane speed (Inline graphic) and the speed of sound is the nondimensional Mach number (Inline graphic) that serves as a criterion for determining the regime of flight as well as the possible occurrence of special phenomena such as shock waves295299. The Mach number is defined as

graphic file with name 41598_2025_7639_Article_Equ67.gif 67

The flight regime with Inline graphic is subsonic, the flight regime with Inline graphic is transonic or sonic, the flight regime with Inline graphic is supersonic, while the condition Inline graphic corresponds to the hypersonic flight regime300304.

The speed of sound for air (as an ideal gas) depends on its absolute temperature (Inline graphic) and on its specific heat ratio (the ratio of specific heat capacities, or the adiabatic index, Inline graphic) as305307

graphic file with name 41598_2025_7639_Article_Equ68.gif 68

In the troposphere layer of the atmosphere (the non-isothermal layer adjacent to the ground), the absolute air temperature is computed according to the lapse rate (Inline graphic K/m) as

graphic file with name 41598_2025_7639_Article_Equ69.gif 69

In the tropopause layer of the atmosphere (the isothermal layer next to the troposphere layer), the absolute air temperature is assumed to have a constant value of 216.65 K. Thus,

graphic file with name 41598_2025_7639_Article_Equ70.gif 70

The specific heat ratio is treated here as a constant with a value of 1.4, which is commonly assigned to ambient air308310.

graphic file with name 41598_2025_7639_Article_Equ71.gif 71

Summary of equations, variables, and constants

In the current section, we provide a summary of the overall six-degree-of-freedom (6-DOF) fluid mechanics problem in terms of the mathematical structure.

Summary of equations

The current fluid mechanics problem is described by a nonlinear differential–algebraic equations (DAE) system, consisting of 35 equations that are either ordinary differential equations (ODE) or algebraic equations. These 35 scalar equations can be grouped into nine categories as shown in Table 5. It should be noted that these groups are not mathematically decoupled; the grouping is based on the scope of use for the equations.

Table 5.

Summary of the 35 flight mechanics equations in the current study.

Equations group Equations count
Body-fixed axes angular velocity components 3
Wind-axes linear-momentum equations (and the dynamic pressure) 4
Body-axes angular-momentum equations (including the auxiliary moments) 6
Inertial velocity components 3
Flight path angles 2
Body-fixed axes aerodynamic forces 3
Body-fixed axes total moments 3
Aerodynamic and stability (moment) coefficients 9
Air density (and the flight altitude) 2
Total 35

Summary of variables

The aforementioned 35 flight mechanics equations have 39 independent variables, which are listed in Table 6 as categorized groups. It should be noted that constant parameters that appear in the differential–algebraic equations (such as the wing planform area and the airplane mass) are not included among the flight variables, because these parameters remain invariant during the entire flight trajectory. In addition, the time derivative of a flight variable is not considered an additional separate variable.

Table 6.

Summary of the 39 flight mechanics variables in the current study.

Variables group Variables type (in InvSim) Variables symbols Variables count
Inertial coordinates and roll Euler angle input Inline graphic 4
Pitch and yaw Euler angles intermediate Inline graphic 2
Body-axes angular velocity components intermediate Inline graphic 3
Wind-axes coordinates for the linear velocity intermediate Inline graphic 3
Spherical angular coordinates (flight path angles) for the airplane’s inertial location intermediate Inline graphic 2
Body-axes aerodynamic forces (and dynamic pressure) intermediate Inline graphic 4
Body-axes total moments (and auxiliary moments) intermediate Inline graphic 6
Aerodynamic and stability coefficients intermediate Inline graphic 9
Air density and flight altitude intermediate Inline graphic 2
Flight controls output Inline graphic 4
Total 39

The difference between the number of equations and the number of variables is four, and this is the number of input constraints that should be specified in order to be able to integrate the flight mechanics system and obtain a unique solution. The four constraints in the case of our inverse simulation (InvSim) flight mechanics model are the temporal profiles of the three inertial coordinates (Inline graphic) and the temporal profile of the roll angle (Inline graphic). The output variables to be obtained by the InvSim flight mechanics solver are the discrete profiles of four airplane controls, which are the thrust and the three deflection angles of the control surfaces (Inline graphic). The remaining 31 variables (such as the angle of attack Inline graphic, and the Euler pitch angle Inline graphic) are intermediate quantities that can evolve over time during the flight maneuver in response to changes in other related variables.

Summary of constants

In addition to the 39 flight variables (that generally vary during the flight maneuver), various constant parameters need to be defined once, and these values remain unchanged during the entire flight simulation. The total number of constants in the proposed InvSim model is 30, which are classified as 29 parameters for describing the airplane and its performance, and one parameter related to the trajectory (the initial altitude). These 30 constant parameters are summarized in Table 7, which are organized as related groups.

Table 7.

Summary of the 30 constant parameters for the presented inverse simulation model.

Parameters group Parameters symbols Parameters count
Airplane mass Inline graphic 1
Wing planform (projected) area Inline graphic 1
Mean aerodynamic chord (MAC) Inline graphic 1
Wing span Inline graphic 1
Mass moments and products of inertia about body axes Inline graphic 6
Aerodynamic-force constants Inline graphic 5
Longitudinal stability derivatives Inline graphic 4
Lateral stability derivatives Inline graphic 5
Directional stability derivatives Inline graphic 5
Initial altitude Inline graphic 1
Total 30

It should be noted that eight physical universal constants are not counted among the 30 parameters, because these are not customizable quantities. These universal constants (for the proposed InvSim algorithm) are

  1. Gravitational acceleration (Inline graphic = 9.81 m/s2)

  2. Tropospheric lapse rate magnitude (Inline graphic = 0.0065 K/m)

  3. Ideal gas constant for air (Inline graphic = 287 J/kg.K)

  4. Standard sea-level air density (Inline graphic = 1.225 kg/m3)

  5. Standard sea-level air absolute temperature (Inline graphic = 288.15 K)

  6. Standard altitude of the troposphere-tropopause transition (11,000 m)

  7. Standard air density at the troposphere-tropopause transition (Inline graphic = 0.3636309 kg/m3)

  8. Standard air absolute temperature within the tropopause layer (Inline graphic = 216.65 K)

    While it is possible to upgrade the model by treating the gravitational acceleration as a function of altitude (in this case, the number of flight variables increases from 39 to 40); the gain from this upgrade is not justified. This change largely increases the mathematical complexity of the model, where the time derivative of the gravitational acceleration appears in the mathematical expressions, while such a derivative is practically zero.

    To demonstrate this (and to justify treating the gravitational acceleration as a universal constant), we first point out that the altitude-dependent gravitational acceleration (Inline graphic) declines with the geometric altitude according to the following quadratic relationship:

    graphic file with name 41598_2025_7639_Article_Equ72.gif 72

    At a geometric altitude of 11,000 m (which is the upper limit of the troposphere layer), the altitude-dependent gravitational acceleration is 99.656% of its approximated constant value (at sea level); and at a higher geometric altitude of 20,000 m (which is the upper limit of the tropopause layer), the altitude-dependent gravitational acceleration drops further to 99.375% its approximated constant value. Thus, the relative drops in the gravitational acceleration at 11,000 m and 20,000 m geometric altitudes are only 0.344% and 0.625%; respectively. As dimensional drops in the altitude-dependent gravitational acceleration from an assigned sea-level value of 9.81 m/s2; the respective drops at 11,000 m and 20,000 m geometric altitudes are 0.0033788 m/s2 (Inline graphic = 9.81 m/s2 – 9.776212 m/s2) and 0.0061303 m/s2 (Inline graphic = 9.81–9.748697 m/s2); respectively. At a rate-of-climb (RoC) of 2000 ft/min (33.33 ft/s, 10.16 m/s, or 36.58 km/h), which is typical for a commercial jet airplane, an altitude of 11,000 m (36,089 ft) can be reached after a continuous climb for about 1083 s (about 18 min); and this means that the average rate of change of the altitude-dependent gravitational acceleration in this case is approximately 3.12 × 10–6 m/s3, which is nearly zero311318. Although high-performance fighter airplanes may achieve much higher climb rates and descent rates, such as 20,000 ft/min (10 times the typical rates for commercial jet airplanes), the time rate of change in the altitude-dependent gravitational acceleration in such cases remains negligible319325.

  9. Specific heat ratio for air (Inline graphic = 1.4)

    Although the specific heat ratio for air (as an ideal non-monatomic gas) is actually a function of its temperature (while it is temperature-independent or monatomic gases like helium and argon326,327), and this air temperature is a function of the altitude within the troposphere layer; treating the specific heat ratio as a universal constant as adopted here is appropriate given the narrow range of temperatures for atmospheric air during an airplane flight328330. The specific heat ratio is the ratio of the specific heat capacity at constant pressure (Inline graphic) to the specific heat capacity at constant volume (Inline graphic)331,332. In this case of a constant specific heat ratio, air is assumed to be a “calorically perfect” gas, which is a special simplified case of ideal gases333. For air, the specific heat ratio is 1.4022 at 288.15 K (15.00 °C) and it increases to 1.4027 at 216.65 K (–56.50 °C); while it drops to 1.4015 at 323.15 K (50.00 °C); and these values show the weak deviation from the nominal used value of 1.4334.

InvSim customized equations of motion

In this part, we present a variant of the general fluid mechanics equations of motion, which are adapted for use in an inverse simulation (InvSim) mode. These customized equations are presented as nine groups in the next nine subsections, with a different order than the one presented earlier for the general fluid mechanics equations (not customized for the inverse mode) in “General equations of motion” section; this change in the order of presentation facilitates explaining the rationale behind the need for additional derivative expressions beyond those appearing in the general flight mechanics equations.

As a summary before starting the detailed formulation, the three unknown control deflection angles (Inline graphic) are computed from three algebraic equations, while the unknown thrust force (Inline graphic) is computed among eight flight variables through a classical fourth-order Runge–Kutta method (RK4) for integrating a system of nonlinear ordinary differential equations (ODE)335340.

InvSim aerodynamic and stability coefficients

The formulas presented earlier in “Aerodynamic and stability coefficients” section for the aerodynamic lift coefficient (Inline graphic), the aerodynamic drag coefficient (Inline graphic), the side-force coefficient (Inline graphic), and the three body-axes aerodynamic coefficients (Inline graphic) remain in use in the transformed InvSim mode of the flight mechanics equations of motion. For completeness, these expressions are repeated here (while being assigned their same original equation numbers).

graphic file with name 41598_2025_7639_Article_Equ73.gif 49
graphic file with name 41598_2025_7639_Article_Equ54.gif 54
graphic file with name 41598_2025_7639_Article_Equ55.gif 55
graphic file with name 41598_2025_7639_Article_Equ56.gif 56
graphic file with name 41598_2025_7639_Article_Equ57.gif 57
graphic file with name 41598_2025_7639_Article_Equ58.gif 58

However, the original expressions for the body-axes total-moment coefficients (Inline graphic) are restructured into three explicit expressions to obtain the three control surface deflection angles. The original expressions for the body-axes total-moment coefficients (Inline graphic) are repeated below to facilitate the derivation of the restructured ones.

graphic file with name 41598_2025_7639_Article_Equ59.gif 59
graphic file with name 41598_2025_7639_Article_Equ60.gif 60
graphic file with name 41598_2025_7639_Article_Equ61.gif 61

From Eq. (60), the restructured explicit expression for the necessary control surface deflection angle for the elevators (Inline graphic) can be obtained as

graphic file with name 41598_2025_7639_Article_Equ82.gif 73

Solving Eqs. (59 and 61) simultaneously for the necessary control surface deflection angle of the ailerons (Inline graphic) and the rudder (Inline graphic) gives

graphic file with name 41598_2025_7639_Article_Equ83.gif 74
graphic file with name 41598_2025_7639_Article_Equ84.gif 75

The coupling between the rolling and yawing moments is noticeable from Eqs. (59 and 61), and these two moments are together decoupled from the pitching moment as indicated by Eq. (60); this behavior is known for fixed-wing airplanes341346.

InvSim three moments

In “Three moments” section of the original general (not customized for the inverse mode) flight mechanics formulation, the total moment vector was resolved into three components (three moments; Inline graphic) along the body axes (Inline graphic); and these body-referenced moments were expressed in terms of three nondimensional moment coefficients (Inline graphic), respectively. However, in the previous “InvSim aerodynamic and stability coefficients” section, we showed that the proposed InvSim flight mechanics formulation requires the values of these nondimensional moment coefficients to algebraically obtain the corresponding control surface deflection angles (Inline graphic).

Therefore, the original expressions for (Inline graphic) are inverted here to be explicit expressions for obtaining (Inline graphic), as follows:

graphic file with name 41598_2025_7639_Article_Equ85.gif 76
graphic file with name 41598_2025_7639_Article_Equ86.gif 77
graphic file with name 41598_2025_7639_Article_Equ87.gif 78

InvSim angular-momentum equations

In the part of the original general flight mechanics formulation covered earlier in “Three moments” section, algebraic expressions for three auxiliary moments (Inline graphic) were given as functions of the body-referenced angular velocity components (Inline graphic) and body-referenced total moments (Inline graphic); with the body-axes inertia components (Inline graphic) being constant geometric parameters. These expressions are repeated below.

graphic file with name 41598_2025_7639_Article_Equ18.gif 18
graphic file with name 41598_2025_7639_Article_Equ19.gif 19
graphic file with name 41598_2025_7639_Article_Equ20.gif 20

Also, the original general flight mechanics formulation in “Three moments” section included three main rotational equations of motion for the airplane about its body axes, which are repeated below.

graphic file with name 41598_2025_7639_Article_Equ21.gif 21
graphic file with name 41598_2025_7639_Article_Equ22.gif 22
graphic file with name 41598_2025_7639_Article_Equ23.gif 23

As explained before, (Inline graphic) is the determinant of the inertia tensor.

However, according to the discussion given in the previous subsection, the body-referenced total moments (Inline graphic) are needed in order to obtain the corresponding nondimensional moment coefficients (Inline graphic). These body-referenced total moments are obtained by solving simultaneously the above system of three equations, Eqs. (2123), for the auxiliary moments (Inline graphic) as three explicit expressions whose right-hand side include the body-referenced angular accelerations (Inline graphic). The resultant explicit symbolic expressions for (Inline graphic) and (Inline graphic) are very large and thus are not shown here; but we designate them by the placeholder symbols (Inline graphic) and (Inline graphic), respectively (the subscript “Inline graphic” here indicates a function that can be simply evaluated to yield a left-hand side variable if all right-hand side elements are known). However, the explicit symbolic expressions for (Inline graphic) is simple enough to be shown here, and it is given below and we designate it by the symbol (Inline graphic) to distinguish it from the original explicit expression in Eq. (20) that also has (Inline graphic) in its left-hand side as a single quantity, but that original expression is not used in the proposed InvSim algorithm presented here, because it is replaced by the derived expression below for (Inline graphic) in Eq. (79).

graphic file with name 41598_2025_7639_Article_Equ94.gif 79

After the values of the auxiliary moments become known at a given time station, the body-referenced total moments (Inline graphic) can be computed using an adapted version of the equations that defined the auxiliary moments (Inline graphic); namely Eqs. (181920), respectively. These adapted equations are suitable for computing the numerical values of (Inline graphic), and they have the following form:

graphic file with name 41598_2025_7639_Article_Equ95.gif 80
graphic file with name 41598_2025_7639_Article_Equ96.gif 81
graphic file with name 41598_2025_7639_Article_Equ97.gif 82

InvSim angular velocity vector in body axes

The expressions relating the body-referenced roll rate (Inline graphic), body-referenced pitch rate (Inline graphic), and body-referenced yaw rate (Inline graphic) to the Euler rates and Euler angles remain the same. These expressions were presented in “Angular velocity vector in body axes” section through Eqs. (24), which are repeated below.

graphic file with name 41598_2025_7639_Article_Equ2.gif 2
graphic file with name 41598_2025_7639_Article_Equ3.gif 3
graphic file with name 41598_2025_7639_Article_Equ4.gif 4

The above algebraic expressions for the body-referenced angular velocities (Inline graphic) are used to find initial values (Inline graphic) for them at the beginning of the maneuver’s numerical simulation, at the time (Inline graphic) = 0.

According to the InvSim customized expressions (Inline graphic) discussed earlier in the previous “InvSim angular-momentum equations” section for the auxiliary moments (Inline graphic), it is required to know the values of the derivatives of the body-referenced angular velocities (Inline graphic) in order to be able to compute the dependent values of the auxiliary moments (Inline graphic). Expressions for these three body-referenced angular accelerations (Inline graphic) are obtained symbolically by direct differentiation of the expressions for (Inline graphic) as presented in Eqs. (24). The resultant expressions for the body-referenced angular accelerations are

graphic file with name 41598_2025_7639_Article_Equ101.gif 83
graphic file with name 41598_2025_7639_Article_Equ102.gif 84
graphic file with name 41598_2025_7639_Article_Equ103.gif 85

InvSim flight path angles

The obtained explicit expressions in Eqs. (8385) for the three body-referenced angular accelerations (Inline graphic) in the previous subsection contain the Euler angular velocities (Inline graphic) and the Euler angular accelerations (Inline graphic). While (Inline graphic) is considered available (either through numerically differentiating twice the input series values of Inline graphic or through evaluating a symbolic expression for (Inline graphic) if (Inline graphic) is provided as a double-differentiable function of time), additional analytical expressions are needed for (Inline graphic) and (Inline graphic). These can be obtained by symbolically differentiating the two algebraic equations that relate the two flight path angles (Inline graphic) to the three Euler angles (Inline graphic) as well the angle of attack (Inline graphic) and the sideslip angle (Inline graphic). These two algebraic equations to be differentiated are repeated below.

graphic file with name 41598_2025_7639_Article_Equ33.gif 33
graphic file with name 41598_2025_7639_Article_Equ34.gif 34

After performing symbolic differentiation twice for the above two algebraic equations, the resultant equations can be manipulated to obtain the sought explicit expressions for (Inline graphic) and (Inline graphic), as well as for the explicit expressions for the first derivatives (Inline graphic) and (Inline graphic). These (Inline graphic) and (Inline graphic) expressions are used in initializing the time loop for the numerical integration through the maneuver’s discrete-time stations by estimating (Inline graphic) and (Inline graphic) at the start of the maneuver’s simulation.

The symbolic explicit expressions for (Inline graphic, Inline graphic, Inline graphic, Inline graphic) are too complicated to be conveniently shown here. So, we alternatively refer to these four symbolic explicit expressions using the placeholder symbols (Inline graphic, Inline graphic, Inline graphic, and Inline graphic); respectively.

InvSim linear-momentum equations and equilibrium

The first equation of the wind-referenced linear-momentum equations, which is Eq. (8), can be reformulated to be an explicit expression for the thrust force (Inline graphic), as follows:

graphic file with name 41598_2025_7639_Article_Equ106.gif 86

This algebraic explicit expression for the thrust (Inline graphic) is used at the InvSim initialization stage for computing an initial thrust value (Inline graphic). Although the previous expression for (Inline graphic) has a singularity at (Inline graphic = ± 90° or ± π/2 rad), these conditions are not realistic for a fixed-wing airplane, as they imply that the airplane is flying relatively perpendicular to its wing’s surface. Also, although the previous expression for (Inline graphic) has a singularity at (Inline graphic = ± 90° or ± π/2 rad), these conditions are not realistic for a fixed-wing airplane as they imply that the airplane is drifting sideways by flying laterally, perpendicular to its longitudinal axis, similar to a cylinder interacting with a fluid flow and is oscillating laterally perpendicular to its longitudinal axis due to its interaction with the surrounding flow347352.

The explicit symbolic expressions (Inline graphic and Inline graphic) discussed in the previous subsection involve the second time derivative of the angle of attack (Inline graphic) and the second time derivative of the sideslip angle (Inline graphic), which appear in the right-hand side of (Inline graphic) and (Inline graphic). Therefore, two additional explicit symbolic expressions (to be denoted by the placeholder symbols Inline graphic and Inline graphic) are needed, which describe mathematically the second derivatives (Inline graphic and Inline graphic), respectively. Like (Inline graphic and Inline graphic), the explicit expressions (Inline graphic and Inline graphic) are very elaborate and are not shown here, but we discuss next how these expressions can be obtained.

We provide below a reformulated version of the third equation of the wind-referenced linear momentum equations, which is Eq. (10) that represents the component along the third wind axis (Inline graphic), as an explicit expression for the first derivative of the angle of attack (Inline graphic).

graphic file with name 41598_2025_7639_Article_Equ107.gif 87

Differentiating the above expression once with respect to time yields the explicit expression (Inline graphic).

A reformulated version of the second equation of the wind-referenced linear momentum equations, which is Eq. (9), representing the linear momentum equation along the lateral wind axis (Inline graphic), is provided below as an explicit expression for the first derivative of the sideslip angle (Inline graphic).

graphic file with name 41598_2025_7639_Article_Equ108.gif 88

Differentiating the above expression once with respect to time yields the explicit expression (Inline graphic).

Although either of the two previous expressions for (Inline graphic and Inline graphic) has a singularity at (Inline graphic = 0), this condition is not realistic for a fixed-wing airplane, as it implies that the airplane is stagnant relative to the ambient air (a hovering condition).

In the explicit expressions (Inline graphic and Inline graphic), the first derivative of the thrust (Inline graphic) appears on the right-hand side. In order to construct a symbolic explicit expression for (Inline graphic); Eq. (86), which is an explicit expression for (Inline graphic) is differentiated once with respect to time. The resultant expression is very elaborate and thus is not shown, but is denoted by the placeholder symbol (Inline graphic).

The discussions given before in “Linear-momentum equations and equilibrium” section and “Aerodynamic and stability coefficients” section regarding the interpretation of the angle of attack (Inline graphic) appearing in the general flight mechanics modeling, and the relation of this angle to the conventional angle of attack (Inline graphic) and its equilibrium value (Inline graphic) remain the same here for the InvSim mode of flight mechanics modeling.

InvSim inertial velocity

Going back to the needed explicit expressions (Inline graphic and Inline graphic) discussed earlier in “InvSim flight path angles” section; the explicit expression (Inline graphic) has the second time derivative of the azimuth flight path angle (Inline graphic) in its right-hand side, while the right-hand side of the explicit expression (Inline graphic) requires the second time derivative of both of the azimuth flight path angle (Inline graphic) and the elevation flight path angle (Inline graphic).

The original expression for the inertial Cartesian velocity components (Inline graphic) in “Inertial velocity” section can be mathematically altered to give two explicit expressions for these flight path angles (Inline graphic and Inline graphic). We recall below the original equations for the ground-referenced velocity components.

graphic file with name 41598_2025_7639_Article_Equ30.gif 30
graphic file with name 41598_2025_7639_Article_Equ31.gif 31
graphic file with name 41598_2025_7639_Article_Equ32.gif 32

From these equations, the velocity of the airplane’s center of gravity can be expressed in terms of the spherical flight-path coordinates (Inline graphic) as

graphic file with name 41598_2025_7639_Article_Equ112.gif 89
graphic file with name 41598_2025_7639_Article_Equ113.gif 90
graphic file with name 41598_2025_7639_Article_Equ114.gif 91

The quantity Inline graphic is the projected velocity component in the horizon plane (Inline graphic), and it is equivalent to (Inline graphic) according to Eqs. (30 and 31). It can be shown that this quantity is also equivalent to (Inline graphic), given that (Inline graphic) and (Inline graphic).

It should also be noted that the derivatives of the inertial Cartesian coordinates (Inline graphic) are computed either numerically using approximated finite difference expressions that are applied to the input discrete series of the inertial Cartesian coordinates (Inline graphic), or through evaluating symbolic expressions if these coordinates are described symbolically as differentiable functions. In either case, these derivatives (Inline graphic) are assumed to eventually become available as a series of discrete values covering the entire maneuver duration.

The numerical values of the second derivatives of the inertial Cartesian coordinates (Inline graphic) are needed to evaluate the symbolic expressions for the first derivatives of the two flight angles (Inline graphic and Inline graphic), while the numerical values of the third derivatives of these inertial Cartesian coordinates (Inline graphic) are needed to evaluate the symbolic expressions for the second derivatives of the two flight angles (Inline graphic and Inline graphic). Again, the values of (Inline graphic) and (Inline graphic) can be obtained through numerical differentiation or symbolic differentiation (followed by numerical substitution at each of the time stations along the maneuver trajectory), depending on how the input ground-referenced coordinates (Inline graphic) for the maneuver trajectory are described.

Similarly, expressions for the first and second derivatives of the velocity magnitude (Inline graphic and Inline graphic) can be developed in terms of the inertial accelerations (Inline graphic) and inertial jerks (Inline graphic). Alternatively, one may compute the discrete values of (Inline graphic) from the numerical values of (Inline graphic) by using the above Eqs. (8991), and then applying finite difference formulas to compute the values of the first time derivatives (Inline graphic) and the second time derivatives (Inline graphic).

We list below standard formulas of the finite difference method (FDM) to approximate a first derivative (Inline graphic), a second derivative (Inline graphic), and a third derivative (Inline graphic), of a generic time-dependent function Inline graphic, if its values at uniformly-spaced time stations having a fixed time step (Inline graphic) are known353357. These formulas are second-order-accurate, which means that the discretization error decays at a rate that is related to the decay rate of the time step squared. For each derivative order (one, two, or three); we provide three types of finite difference formulas, namely (1) forward difference (FD), (2) central difference (CD), and (3) backward difference (BD). At the initial time station (assigned a time station index Inline graphic = 1), forward difference should be used, since no past “backward” values are available. At the final time station (assigned a time station index Inline graphic), backward difference should be used, since no future “forward” values are available. At other intermediate time stations (having time station index values of Inline graphic for Inline graphic and Inline graphic; but Inline graphic for Inline graphic), central difference is preferred because it involves less computation and it is symmetric (considering both past and future data). The finite difference formulas are listed in Table 8. In this table, the subscript indices (such as “Inline graphic”, “Inline graphic”, and “Inline graphic”) refer to the time position relative to the generic time station (Inline graphic).

Table 8.

General finite difference formulas for numerical differentiation.

Derivative Difference type Expression
Inline graphic Forward Inline graphic
Inline graphic Central Inline graphic
Inline graphic Backward Inline graphic
Inline graphic Forward Inline graphic
Inline graphic Central Inline graphic
Inline graphic Backward Inline graphic
Inline graphic Forward Inline graphic
Inline graphic Central Inline graphic
Inline graphic Backward Inline graphic

With this, the numerical values of the following 12 quantities should be known at each time station before the main InvSim flight mechanics simulation starts:

  1. Inline graphic

  2. Inline graphic

  3. Inline graphic

  4. Inline graphic

We provide below symbolic expressions for the first derivatives (Inline graphic), while the symbolic expressions for the second derivatives (Inline graphic) are much more detailed and thus are not provided here, but we denote them by the placeholder symbols (Inline graphic), respectively.

graphic file with name 41598_2025_7639_Article_Equ115.gif 92
graphic file with name 41598_2025_7639_Article_Equ116.gif 93
graphic file with name 41598_2025_7639_Article_Equ117.gif 94

Either of the two previous expressions for (Inline graphic) and (Inline graphic) has a singularity at (Inline graphic), which corresponds to (Inline graphic = ± 90° or ± π/2 rad). In a real flight setting, this means that the airplane is flying straight up or straight down with respect to the fixed ground (thus, the airplane is flying perpendicular to the horizon). This is a restriction in the present InvSim numerical algorithm, which fails in this exceptional case. Therefore, such two particular flight movements (vertical ascent, nose-up; and vertical descent, nose-down) should be excluded, although some military airplanes are capable of performing such unconventional flight situations, with their thrust exceeding their weight358360. For civil airplanes and military airplanes not performing this extreme maneuver, this restriction is not a concern.

InvSim three aerodynamic forces

In the proposed InvSim here for the inverse simulation of flight mechanics problems, the nondimensional aerodynamic coefficients along the body axes (Inline graphic) are used in lieu of the dimensional aerodynamic forces along the body axes (Inline graphic). Therefore, the explicit expressions for the aerodynamic forces presented in the original flight mechanics formulation, Eqs. (3537) in “Three aerodynamic forces” section, are not needed for performing the InvSim computations. However, these expressions can still be used for computing and reporting these aerodynamic forces as supplementary post-processing quantities.

This optionality is similar to the one discussed earlier in subsection 3.9 for auxiliary quantities such as the speed of sound (Inline graphic).

InvSim air density and speed of sound

The variation of the air density (Inline graphic), air temperature (Inline graphic), and speed of sound (Inline graphic) with the altitude (Inline graphic) as described by the International Standard Atmosphere (ISA) model and discussed earlier for the general flight mechanics formulation are independent of the mode of solving the flight mechanics problem (either forward mode or inverse mode). Rather, the dependence of these three air properties on the altitude is described by standalone relationships that serve to supply the proper air conditions at the altitude of each time station.

Therefore, this part of the InvSim algorithm is not repeated here, as it is the same as the one discussed earlier for the general flight mechanics formulation in “Air density and speed of sound” section.

We clarify again that only the air density is needed in the InvSim algorithm computations, while the air temperature only facilitates computing the local speed of sound in air, which in turn permits computing the Mach number as a supplementary flight feature.

InvSim numerical algorithm

In this section, we describe the main stages of the numerical integration algorithm for solving the inverse simulation (InvSim) flight mechanics problem for an airplane, based on the miscellaneous equations and theoretical overview given before.

In summary, the algorithm has three main stages, namely (1) pre-processing of inputs, (2) initialization, and (3) time loop Runge–Kutta method.

Pre-processing of Inputs

This first stage of the InvSim numerical algorithm can be further divided into the following 17 steps:

  1. The initial time for the trajectory is set to (Inline graphic = 0). If the entire trajectory duration is (Inline graphic), then the total number of time stations is
    graphic file with name 41598_2025_7639_Article_Equ118.gif 95
    where (Inline graphic) is the uniform time step; and the number of time steps (the transitions between successive time stations) is (Inline graphic) or (Inline graphic).
  2. The 30 constants needed for defining the airplane’s geometry and its aerodynamic/stability behavior, as well as the initial altitude (as listed in “Summary of constants” section), are received by the user.

  3. The four main InvSim inputs are also received by the user, either as analytical (symbolic) expressions or as equally-spaced discrete values (a quantitative vector) with a constant time step (Inline graphic). These four main inputs to the InvSim algorithm are
    • Ground-referenced inertial coordinates of the maneuver trajectory: Inline graphic
    • roll angle: Inline graphic
  4. Using the initial altitude (Inline graphic) and the values of (Inline graphic) into Eq. (1), all numerical values of the flight altitude (Inline graphic) are obtained.

  5. Using the obtained values of the altitude (Inline graphic) into either Eq. (62) for the troposphere layer or Eq. (66) for the tropopause layer, all numerical values of the air density (Inline graphic) are obtained.

    Optionally, all numerical values of the speed of sound can also be computed as we explained in “Air density and speed of sound” section.

  6. Using either analytical (symbolic) differentiation (if the roll angle Inline graphic is provided in a functional form) or second-order finite difference formulas (if the roll angle Inline graphic is provided as a vector of discrete values), all numerical values of the first derivative of the roll angle (Inline graphic) and the second derivative of the roll angle (Inline graphic) are obtained.

  7. Using either analytical (symbolic) differentiation (if the inertial coordinates Inline graphic are provided in a functional form) or second-order finite difference formulas (if the inertial coordinates Inline graphic are provided as vectors of discrete values); all numerical values of their first derivatives (Inline graphic), second derivatives (Inline graphic), and third derivatives (Inline graphic) are obtained.

  8. Using Eq. (89), all numerical values of the velocity magnitude (Inline graphic) are obtained.

  9. Using the obtained values of the air density (Inline graphic) and the obtained values of the airplane speed (Inline graphic) in Eq. (11), all numerical values of the dynamic pressure (Inline graphic) are obtained.

  10. Using Eq. (90), all numerical values of the azimuth flight path angle (Inline graphic) are obtained.

  11. Using Eq. (91), all numerical values of the elevation flight path angle (Inline graphic) are obtained.

  12. Using second-order finite difference formulas discussed in “InvSim inertial velocity” section, all numerical values of the first derivative (Inline graphic) are obtained.

    Alternatively, Eq. (92) can be used.

  13. Using the Eq. (93), all numerical values of the first derivative (Inline graphic) are obtained.

  14. Using the Eq. (94), all numerical values of the first derivative (Inline graphic) are obtained.

  15. Using second-order finite difference formulas discussed in “InvSim inertial velocity” section, all numerical values of the second derivative (Inline graphic) are obtained.

    Alternatively, the explicit expression (Inline graphic) discussed in “InvSim inertial velocity” section, can be used.

  16. Using the explicit expression (Inline graphic) discussed in “InvSim inertial velocity” section, all numerical values of the second derivative (Inline graphic) are obtained.

  17. Using the explicit expression (Inline graphic) discussed in “InvSim inertial velocity” section, all numerical values of the second derivative (Inline graphic) are obtained.

At the end of this first stage of the InvSim numerical algorithm, the following 27 quantitative vectors (one-dimensional arrays, or series of numerical values) become available, covering all the (Inline graphic) time stations of the flight maneuver:

  1. Inline graphic (the four InvSim inputs)

  2. Inline graphic

  3. Inline graphic

  4. Inline graphic

  5. Inline graphic

  6. Inline graphic

  7. Inline graphic

  8. Inline graphic

    Optionally, four more pre-processing quantitative vectors can be computed entirely from the user’s provided data, before solving for other flight variables; although these four quantities are not core elements within the InvSim algorithm, thus the simulation can be completed without knowing these quantities. These four optional pre-processing vectors are

  9. Inline graphic

Initialization

The initial conditions correspond to the first time station, with the index (Inline graphic = 1). The following initial conditions are implemented:

  1. The initial angle of attack (Inline graphic) and the initial sideslip angle (Inline graphic) are set to zero, in alignment with the equilibrium condition.

  2. The above-mentioned initial zero values for (Inline graphic and Inline graphic) dictate that the initial values of the two unspecified Euler angles (Inline graphic and Inline graphic) take the same initial values of the known spherical flight path angles (Inline graphic and Inline graphic), respectively. Therefore, Inline graphic and Inline graphic. These initial conditions are implied by Eqs. (33) and (34), respectively, as discussed in “Flight path angles” section.

  3. The initial thrust force (Inline graphic) is computed using Eq. (86).

  4. The initial angle of attack rate (Inline graphic) and the initial sideslip angle rate (Inline graphic) are set to zero.

  5. The explicit expressions (Inline graphic and Inline graphic) discussed in “InvSim flight path angles” section imply that the initial Euler yaw rate (Inline graphic) and the initial Euler pitch rate (Inline graphic), respectively, have zero values.

  6. Equations (2, 3, 4) imply that the initial body-referenced angular velocities (Inline graphic) respectively, have zero values.

  7. Equations (83, 84, 85) imply that the initial body-referenced angular acceleration (Inline graphic), respectively, have zero values.

  8. The initial control surface deflection angles (Inline graphic) are computed using Eqs. (737475), respectively.

Time loop Runge–Kutta method

This third stage of the InvSim numerical algorithm is the main stage, where the classical explicit fourth-order Runge–Kutta (RK4) four-step integration method is used to numerically compute the evolution of eight flight variables through a system of nonlinear coupled ordinary differential equations (ODE)361365. These eight variables are:

  • Inline graphic

  • Inline graphic

  • Inline graphic

  • Inline graphic

In this subsection, we describe how these eight variables are advanced in discrete time through the proposed InvSim numerical algorithm from an arbitrary time station (discrete time index Inline graphic) with known values for these eight variables to the next time station (discrete time index Inline graphic) with unknown values for these eight variables.

In the beginning, it might be useful to make a brief description of the classical fourth-order Runge–Kutta method for a generic nonlinear scalar ordinary differential equation (ODE) of the form

graphic file with name 41598_2025_7639_Article_Equ119.gif 96

with a known starting condition denoted as Inline graphic.

The procedure of numerically integrating this exemplar ODE in order to compute the value Inline graphic is as follows:

graphic file with name 41598_2025_7639_Article_Equ120.gif 97

where the superscript asterisk in (Inline graphic, Inline graphic, and Inline graphic) refers to a temporary (intermediate) value.

Considering the proposed InvSim algorithm presented here, the above procedure is to be applied four times per time step as follows:

  1. Either of the four intermediate derivatives for the thrust (Inline graphic) is computed using the explicit expression (Inline graphic) discussed in “InvSim linear-momentum equations and equilibrium” section.

  2. Using the explicit expressions (Inline graphic and Inline graphic) along with Eqs. (87 and 88) discussed in “InvSim linear-momentum equations and equilibrium” section, the intermediate derivatives for the angle of attack (Inline graphic) and the sideslip angle (Inline graphic) are computed, respectively as
    graphic file with name 41598_2025_7639_Article_Eque.gif
    graphic file with name 41598_2025_7639_Article_Equf.gif
    graphic file with name 41598_2025_7639_Article_Equg.gif
    graphic file with name 41598_2025_7639_Article_Equ121.gif 98
  3. Using the explicit expressions (Inline graphic and Inline graphic) along with the explicit expressions (Inline graphic and Inline graphic) discussed in “InvSim flight path angles” section, the intermediate derivatives (Inline graphic) and (Inline graphic) are computed as
    graphic file with name 41598_2025_7639_Article_Equ122.gif 99
  4. Either of the four intermediate derivatives of the body-referenced roll rate (Inline graphic) is computed using Eq. (83). Similarly, either of the four intermediate derivatives of the body-referenced pitch rate (Inline graphic) is computed using Eq. (84), and either of the four intermediate derivatives of the body-referenced yaw rate (Inline graphic) is computed using Eq. (85).

  5. The computed 32 temporary derivatives (eight temporary derivatives for eight variables are computed in each of the four steps of the RK4 procedure) are used to update the eight flight variables of the RK4 procedure at time station (Inline graphic), as follows:
    graphic file with name 41598_2025_7639_Article_Equ123.gif 100

    With this update, one of the sought four output flight controls (namely the thrust, Inline graphic) becomes known at the new time station (Inline graphic).

    We need to obtain the remaining three sought output flight controls (the three moving surface deflection angles; Inline graphic, Inline graphic, Inline graphic). This is done in the remaining part of the InvSim numerical algorithm as explained next.

  6. The new values (at time station Inline graphic) of the eight updated flight variables through the RK4 procedure are used to evaluate several derivative expressions, ending with the body-referenced angular acceleration (Inline graphic) at the new time station (Inline graphic) using Eqs. (838485), respectively.

  7. The obtained new body-referenced angular accelerations (Inline graphic) are used in the explicit algebraic expressions (Inline graphic) to compute the auxiliary moments (Inline graphic) at the new time station (Inline graphic).

  8. The obtained new auxiliary moments (Inline graphic) are used in Eqs. (808182), respectively, to find the corresponding total body-referenced moments (Inline graphic) at the new time station (Inline graphic).

  9. The obtained new total body-referenced moments (Inline graphic) are used in Eqs. (767778), respectively, to find the corresponding nondimensional moment coefficients (Inline graphic) at the new time station (Inline graphic).

  10. Finally, the obtained new nondimensional moment coefficient (Inline graphic) is used in Eq. (73) to find the necessary deflection angle for the elevators (Inline graphic) at the new time station (Inline graphic). Similarly, the obtained new nondimensional moment coefficients (Inline graphic) are used in Eqs. (7475), respectively, to find the necessary deflection angle for the ailerons (Inline graphic) and the necessary deflection for the rudder (Inline graphic) at the new time station (Inline graphic).

    If this is not the last time station (Inline graphic), the above 10 computational actions are repeated to find the four necessary flight control variables (Inline graphic) at the subsequent time station. The latest new values obtained at (Inline graphic) become old “known” values when advancing from the “known” time station (Inline graphic) to the “unknown” (Inline graphic).

InvSim example for Mirage III

Although the main contribution of this work is the detailed presentation of the mathematical formulation and the numerical algorithm for inverse simulation (InvSim) flight mechanics, where the required flight controls corresponding to a desired flight trajectory for a given fixed-wing airplane may be computed; it is a valuable addition to this study to provide an example case in which the proposed InvSim algorithm is applied. This demonstrative example is performed in this section for a set of airplane data representing the family of military fighters called “Mirage III”.

About Mirage III

Mirage III is a family of military all-weather aircraft capable of interception at supersonic speeds and capable of taking off using compact runways. Mirage III airplanes are produced by the French aerospace company Dassault Aviation, which is a major partner of the French national defense366370. The Mirage III family is characterized by a delta wing (shaped as a triangle), and a single powerful jet engine able to deliver a thrust force of about 80 kN (80,000 N).

Figure 9 is a photo taken for the model (V 01) of Mirage III, which provides a general overview of its geometric design371.

Fig. 9.

Fig. 9

Photo of Mirage III (model V 01) in flight. Permission for use was requested from the manufacturer Dassault Aviation.

Although there are different models of Mirage III, we here use the following set of geometric and performance parameters listed in Table 9, which are considered to be an archetype representation of Mirage III. Due to the delta-shaped wing, the characteristic longitudinal length (Inline graphic) and the characteristic lateral length (Inline graphic) are assigned equal values.

Table 9.

Adopted geometric parameters for Mirage III.

Serial number Parameter Value
1 Mass (Inline graphic) 7,400 kg
2 Wing planform area (Inline graphic) 36 m2
3 Mean aerodynamic chord (Inline graphic) 5.25 m
4 Wing span (Inline graphic) 5.25 m
5 Moment of inertia about Inline graphic (Inline graphic) 90,000 kg.m2
6 Moment of inertia about Inline graphic (Inline graphic) 54,000 kg.m2
7 Moment of inertia about Inline graphic (Inline graphic) 60,000 kg.m2
8 Moment of inertia in Inline graphic (Inline graphic) 0 kg.m2
9 Moment of inertia in Inline graphic (Inline graphic) 1,800 kg.m2
10 Moment of inertia in Inline graphic (Inline graphic) 0 kg.m2
11 Lift coefficient at zero conventional angle of attack (Inline graphic) 0
12 Slope of lift coefficient (Inline graphic) 2.204 1/rad
13 Drag coefficient at zero lift (Inline graphic) 0.015
14 Drag polar parameter (Inline graphic) 0.4
15 Side-force coefficient parameter (Inline graphic) –0.6 1/rad
16 Longitudinal stability parameter (Inline graphic) 0
17 Longitudinal stability parameter (Inline graphic) –0.17
18 Longitudinal stability parameter (Inline graphic) –0.4
19 Longitudinal stability parameter (Inline graphic) –0.45 1/rad
20 Lateral stability parameter (Inline graphic) –0.05 1/rad
21 Lateral stability parameter (Inline graphic) –0.25
22 Lateral stability parameter (Inline graphic) 0.06
23 Lateral stability parameter (Inline graphic) –0.3
24 Lateral stability parameter (Inline graphic) 0.018 1/rad
25 Directional stability parameter (Inline graphic) 0.15 1/rad
26 Directional stability parameter (Inline graphic) 0.055
27 Directional stability parameter (Inline graphic) –0.7
28 Directional stability parameter (Inline graphic) 0
29 Directional stability parameter (Inline graphic) –0.085 1/rad

The above value of the wing area is gross, thus it includes part of the projected fuselage area located between the wing tips. The equality of the span and chord in the above table is an unusual feature of the tailless Mirage fighter airplanes, given their delta wing having a steep rearward sweep angle and a low aspect ratio near unity.

In addition, we select an initial altitude of Inline graphic = 0. This means that the trajectory is assumed to start from a point located at the mean sea level, and this is a reasonable assumption, and it helps in standardizing our proposed maneuver as a numerical testing case in flight mechanics.

Proposed test maneuver

The flight maneuver we propose here to verify the applicability of the presented inverse simulation (InvSim) numerical algorithm is a simple one, being a horizontal north-wise straight flight at a constant altitude of 5,000 m and with a constant flight speed of 150 m/s while performing a continuous double-roll rotation with a non-linear angular profile constituting of a two-harmonic function. This simple maneuver has several advantages, making it a favored test not only in the current study but also in general for the flight mechanics modeling community globally, including those who want to develop the presented InvSim algorithm, and test its implementation.

With this proposed test maneuver, it is easier to inspect the computational procedure of various flight variables than with a complex maneuver. In our proposed maneuver, different vector quantities reduce practically to an array of a single value repeated over all time stations. In addition, various derivatives become zero throughout the entire maneuver. In addition, all the main InvSim input profiles can be specified through concise analytical (symbolic) functions, rather than as extensive numerical lists.

The four main inputs to the InvSim flight mechanics algorithm in our test maneuver are

graphic file with name 41598_2025_7639_Article_Equ124.gif 101

where (Inline graphic) are in meters, and (Inline graphic) is in radians.

Taking the derivatives of the roll angle (Inline graphic) gives the following expressions for its Euler rate (Inline graphic), and Euler acceleration (Inline graphic):

graphic file with name 41598_2025_7639_Article_Equ125.gif 102

Figures 10, 11, and 12 illustrate the temporal profile of the roll angle (Inline graphic), its first-time derivative (Inline graphic), and its second-time derivative (Inline graphic); respectively, during the proposed double-roll maneuver. In these figures, the angles and their gradients are expressed in degrees, degrees per second, or degrees per square second (through scaling by the multiplicative value Inline graphic = 57.296°/rad) for easier comprehension since the degree unit is broadly used. The roll angle changes from 0° to 720° over a duration of 30 s. The rate (Inline graphic) has a non-negative symmetric profile, peaking at the midpoint of 15 s to reach a maximum value of 56.549°/s; while (Inline graphic) has an antisymmetric profile, with a positive peak of 6.838°/s2 at 9.123 s and a negative peak of − 6.838°/s2 at 20.877 s.

Fig. 10.

Fig. 10

Profile of the roll angle in the test maneuver.

Fig. 11.

Fig. 11

Profile of the first time derivative of the roll angle in the test maneuver.

Fig. 12.

Fig. 12

Profile of the second time derivative of the roll angle in the test maneuver.

Table 10 lists some of the computed characteristics for the test maneuver. These characteristics aid in validating any developed computational implementations by other researchers because these quantities can be computed at the initial time station (Inline graphic = 1, Inline graphic = 0) independently of the time loop calculations. In addition, these quantitative characteristics provide useful insights into the nature of this numerical test maneuver when viewed as a real flight mission. It should be noted that due to the maneuver’s simplicity, these scalar quantities are actually vectors of constants (vectors or arrays representing frozen or static variables), thus all the time (Inline graphic) stations have the same value for each characteristic quantity.

Table 10.

Characteristics of the proposed test maneuver of the Mirage III fighter airplane.

Serial number Fixed-value variable Value
1 Flight speed (Inline graphic) 150 m/s (540 km/h, 291.577 knots)
2 Azimuth flight path angle (Inline graphic)
3 Elevation flight path angle (Inline graphic)
4 Altitude (Inline graphic) 5,000 m (16,404 ft; 3.1069 mi)
5 Air density (Inline graphic) 0.73587 kg/m3
6 Dynamic pressure (Inline graphic) 8,278.56 Pa (0.08170 atm)
7 Air absolute temperature (Inline graphic) 255.65 K (–17.50 °C)
8 Air absolute pressure (Inline graphic) 53,992 Pa (0.53286 atm)
9 Speed of sound in air (Inline graphic) 320.50 m/s (1,153.8 km/h ; 623.00 knots)
10 Mach number (Inline graphic) 0.4680
11 Equilibrium conventional angle of attack (Inline graphic) 6.3322°

Inverse simulation results

We implemented the proposed InvSim numerical algorithm as a MATLAB/Octave computer code, and we applied it here using GNU Octave version 6.1.0372. The algorithm does not require any specialized toolboxes, which is a big advantage of simplicity. The time step (Inline graphic) we used is 0.001 s (thus, there is a total of Inline graphic = 30,001 time stations). We tried different values for the time step, and we found that this selected value is proper. With much lower values (such as 0.01 s), the simulation can still be performed without instability, but minor spurious oscillations or non-smooth variations near sharp peaks may occur.

We start the results of the InvSim algorithm for the test maneuver with the four outputs, which are the flight controls. Figure 13 shows the computed thrust force, which is one of the four flight controls. The equilibrium value is 11,543 N (11.543 kN), which changes between four minima and three maxima until the same equilibrium value is restored at the end of the maneuver. The three encountered maxima values of the thrust are 11,332 N at 11.613 s; 11,535 N at 15.002 s; and 11,348 N at 18.390 s. All these peaks are less than (but close to) the equilibrium value. Knowing the global maximum thrust needed during a prospective maneuver is particularly important because it helps in deciding whether that target maneuver is achievable or not, based on the available maximum thrust for the airplane of concern. The four thrust minima are nearly 4,900 N. Knowing the global minimum thrust needed during a maneuver is also important, because negative values (reversed backward thrust) typically imply that the maneuver is not achievable.

Fig. 13.

Fig. 13

Computed temporal profile of the thrust flight control during the test maneuver.

Figure 14 shows the necessary variations in the rudder control surface deflection angle (Inline graphic) during the test maneuver. The maximum absolute deflection angle here is 40.68°. This value is apparently acceptable in terms of geometric constraints; whereas deflections near or exceeding 60° may not be realistic, and thus suggest that the maneuver is too challenging to be accomplished with the present configuration of the airplane373375. The necessary deflection angles of the elevators (Inline graphic) and the ailerons (Inline graphic) are much smaller than those demanded by the rudder, as shown in Fig. 15. The elevators deflection angle is positive most of the time, with negative values encountered for a brief period near the middle of the maneuver. None of the rudder deflection angle and the ailerons deflection angle are exactly symmetric about the horizontal zero line. The rudder deflection angle is positively biased, with a positive mean value of 1.294°; while the ailerons deflection angle is negatively biased, with a negative mean value of –0.624°.

Fig. 14.

Fig. 14

Computed temporal profile of the rudder flight control during the test maneuver.

Fig. 15.

Fig. 15

Computed temporal profiles of the elevators and ailerons flight controls during the test maneuver.

Figure 16 provides additional information about the inversely simulated test maneuver, through displaying the variations in the angle of attack (and its conventional counterpart) and the sideslip angle. A strong correlation can be qualitatively observed between the sideslip angle and the rudder deflection angle. This can be explained by noting that a positive sideslip angle (when the thrust vector is collinear with the longitudinal body axis) leads to a drift in the airplane’s travel path toward the left (the port side) due to a continuously applied port-wise force component of the thrust, and therefore the path becomes curved. To counteract this and maintain a straight flight path, as in the target test maneuver here, a restoring moment needs to be induced through a positive rudder deflection angle (the rudder tilts toward the starboard/right side). The profile of the conventional angle of attack (Inline graphic) is just a shifted version of the simulation-reported angle of attack (Inline graphic); with the difference being the frozen equilibrium conventional angle of attack (Inline graphic), which is 6.3322° (0.110518 rad) in the current test maneuver. During the maneuver, the conventional angle of attack (Inline graphic) is bound between –6.1129° and 6.3322°; which is a narrow range around the zero value, thus validating the assumption of non-stall.

Fig. 16.

Fig. 16

Computed temporal profiles of the angle of attack and sideslip angle during the test maneuver.

Figure 17 provides the profiles of the two Euler angles that were not specified as input constraints, but were computed by the InvSim algorithm, and these angles are the pitch angle and the yaw (or heading) angle. Figure 18 visualizes the variation of these two airplane attitude angles, but as an orbit plot, where the pitch angle is plotted against the yaw angle with the time being a parametric variable376,377. This particular figure allows for judging the convergence of the simulation as the time step is successively refined. At a satisfactorily small time step (like the one used here, 0.001 s), the orbit plot shows a smooth continuous closed double-loop resembling a cardioid. In fact, this plot shows two double-loop curves on top of each other, each one is formed due to one rolling revolution by the airplane during one-half of the total maneuver duration. At an improper coarser time step (such as 0.01 s, as shown in Fig. 19; and such as 0.02 s, as shown in Fig. 20), these two curves become detached and can be visually differentiated. This means that the profiles of these two Euler angles during the first rolling revolution deviate from their profiles during the second rolling revolution, due to pronounced numerical errors.

Fig. 17.

Fig. 17

Computed temporal profiles of the pitch and yaw angles during the test maneuver.

Fig. 18.

Fig. 18

Computed orbit plot of the pitch and yaw angles during the test maneuver at the used time step of 0.001 s.

Fig. 19.

Fig. 19

Computed orbit plot of the pitch and yaw angles during the test maneuver at a coarse time step of 0.01 s.

Fig. 20.

Fig. 20

Computed orbit plot of the pitch and yaw angles during the test maneuver at a coarser time step of 0.02 s.

Discussion

In this discussion, we would like to make four supplementary comments regarding the presented study.

Contributions

First, the contributions of the study include:

  1. The detailed mathematical formulation of the three-dimensional airplane flight dynamics (the version we presented in our earlier work, which simultaneously and efficiently utilizes four different axes/coordinate systems), and its transformation from the general form to the inverse simulation (InvSim) form: This contribution included more than 80 equations, which were not simply taken from an external source. Instead, beneficial derivations and explanations were provided with the aid of visualizing sketches to properly illustrate the definition of various flight variables and the relationship among some of them. The use of manual and software-based mathematical manipulation is a valuable element of that contribution, and readers unfamiliar with such tools may find this study valuable to them in terms of either introducing them to such tools, or providing them with the final intricate expressions we obtained, making them ready to use.

  2. The detailed numerical algorithm that converts the mathematical InvSim formulation into a clear procedure that can be implemented using computer programming: A large number of details, even small ones such as the quantification of errors in the gravitational acceleration at different altitudes, were presented sufficiently clear such that readers having no prior background these diverse concepts (aeronautics, atmospheric modeling, gas dynamics, thermodynamics) can deal with them without the need to go to other resources. The current study is considered self-contained to a large extent. Furthermore, many references were deliberately selected and used in this study, both as supporting points of reference and as recommended resources for interested readers to use in case they want to gain additional knowledge about a specific topic among those covered here. Even some parts of the study itself were repeated when discussed again, which effectively boosts the flow of the content.

  3. The proposed Mirage III maneuver: This test case for flight mechanics is supplied with enough parameters that allow others to simulate the flight mission. This selected maneuver has several advantages, making it a suitable candidate for adoption as a global numerical experiment to assess other flight mechanics methods. The results of the test case presented here can also serve as a guiding benchmarking solution to compare with.

  4. The detailed mathematical formulation provided here, along with closed-form expressions in multiple axes systems aided by carefully-designed illustrative sketches make this study a valuable educational asset as well for students and educators in undergraduate studies; where the study may be used partly or fully as a course content in different subjects, such as dynamics, flight mechanics, numerical methods, mathematical modeling, differential equations, robotics and artificial intelligence, and aerodynamics378382.

Limitations

Second, it is admitted that the study has some limitations. We try to provide some justification for two of them below. In addition, each limitation listed below can be viewed as a potential extension in the form of another study devoted to addressing it.

  1. Perhaps the main concern about this study is the lack of validation with experimental or other computational results. Although we admit this, and we feel that such validation can be a valuable addition to the study, we attempt here to mitigate this matter through the lack of such comparison data for the specific maneuver selected here (which is a novel test case). Due to the scale and nature of the test problem, being a full-scale fighter airplane, having experimental data is almost impossible. Also, the computed results provided here were supplemented with supporting elements, such as the time-sensitivity analysis and the confirmation that the solution converges successfully as the time step is reduced, the absence of numerical instability, the smoothness of the solution, and the logical pattern of the solution being a double-roll steady-level flight. In addition, the lack of validation with external sources does not affect the contributions made in this study through the mathematical and numerical analysis, the derivation of various equations, procedural steps listed toward solving the InvSim problem, and the exact list of the involved equations, variables, and constants needed. This limitation may be addressed as a separate study.

  2. The airplane configuration considered here neglects some features that can be present in modern airplanes, such as the retractable landing gear. Although these elements do affect the dynamics of the airplane, the simplification made in the presented InvSim model is considered a favorable attribute, making it easier to understand and implement. The model presented can be viewed as a base model, which can be customized and expanded to handle such extra components. Again, this matter may induce additional studies.

  3. The algorithm presented here adopts a north-east-down (NDE) inertial non-geocentric coordinate system for describing the trajectory of the airplane. Alternatively, a non-inertial geocentric earth-centered-earth-fixed (ECEF) coordinate system may be used, but along with another earth-centered inertial (ECI) reference frame383386. Therefore, the NDE choice is more straightforward. It is admitted that the adopted NDE choice and the governing equations presented do not account for the earth’s curvature and rotation. This is another limitation of the presented algorithm. However, to restrain the complexity of the computation and enable the closed-form analytical expressions, this assumption is retained. The extension to accommodate these features may be conducted in separate future studies by the interested readers. To mitigate these missing features, the algorithm should be used for modeling flight missions with a limited range (such as 1,000 km; 620 statute miles; or 870 nautical miles), in which situations the influence of earth’s curvature and rotation can be reasonably dropped.

Comparison with a state-space control model

Third, it might be useful to contrast the InvSim algorithm presented here with a control-based algorithm that follows a typical time-domain state-space representation according to the linear control theory.

For a general dynamic system (including an airplane), the governing equations may for formulated in the following standardized generic form of an ordinary differential system387:

graphic file with name 41598_2025_7639_Article_Equ126.gif 103

where (Inline graphic) is a vector of state variables (states)388 that describe the dynamics of the system and whose current values depend on their previous values, (Inline graphic) is the time derivative of the states, (Inline graphic) is a vector of inputs that is specified externally to steer and influence the system, (Inline graphic) is the time, and (Inline graphic) is a nonlinear function.

An output vector (Inline graphic) can be derived from the states and the inputs through an algebraic relation that takes the following standardized form:

graphic file with name 41598_2025_7639_Article_Equ127.gif 104

where (Inline graphic) is another nonlinear function.

If the two above equations can be linearized such that the original dynamic system is approximated as a linear time invariant (LTI) system389 with no nonlinear terms and without explicit dependence on the time, then these two vector nonlinear equations can be transformed into two vector linear equations as

graphic file with name 41598_2025_7639_Article_Equ128.gif 105
graphic file with name 41598_2025_7639_Article_Equ129.gif 106

where Inline graphic, Inline graphic, Inline graphic, and Inline graphic are matrices that are obtained by a linearization technique. While the elements of these matrices are numerical values, they may either remain constant (corresponding to a common equilibrium or trim condition) or be updated over time in a discrete fashion.

In relation to our study of airplanes with six degrees of freedom, the states can be flight variables (such as the velocity components and the angular velocity components), and the input vector can be a subset of the flight variables. However, the aforementioned LTI approach to the flight mechanics problem is different from the one we presented here in several ways. The InvSim approach we presented aims to solve for the control variables, rather than having them specified as known inputs (as in the LTI approach). Also, the InvSim approach we presented does not require linearization (as in the LTI approach). Instead, it retains the full nonlinearity of the original airplane’s equations of motion. Furthermore, the InvSim approach we presented does not involve any matrix operations, and does not require any matrix–vector multiplication (as in the LTI approach). Therefore, the two approaches are different, and neither one can replace the other.

Quaternion notation versus Euler angles for attitude representation

Fourth, we elaborate on the singularity problem that occurs at ± 90° pitch angle if the orientation of the airplane (or another rigid body) is related to the inertial axes system through the sequence of three Euler angles. When the airplane nose (and its body-fixed longitudinal axis) is pointing vertically up or down, parallel to the inertial gravity axis (the pitch angle thus is Inline graphic =  ± 90°), then a singularity occurs, and it is sometimes referred to as a (gimbal lock) due to the analogy with a gyroscope mounted on a mechanical gimbal system in the real world390,391. This singularity generally occurs when the middle rotation angle in a three-angle sequence takes a particular value that makes the rotation axes of the first and third rotations parallel. When the first and third rotational axes become aligned at this singularity condition, they coincide as a combined rotational axis, and the rotation about this common axis becomes effectively due to the sum of attempted rotations about the first and third axes. In our study, this means that when the pitch angle is ± 90°, the roll and yaw are no longer distinguishable. An attempt to perform a rotation of 2° roll with a rotation of –3° yaw is not distinguishable from an attempt to perform a rotation of –2° roll with a rotation of 1° yaw, since the sum of both rotation angles is the same for both cases (–1°). This singularity is mathematically manifested through the appearance of infinite terms due to having sin(Inline graphic) in the denominator; particularly tan(Inline graphic) and sec(Inline graphic). For example, the following transformation equations fail at this singularity condition392:

graphic file with name 41598_2025_7639_Article_Equ130.gif 107
graphic file with name 41598_2025_7639_Article_Equ131.gif 108

To cope with this singularity issue, the algorithm presented here uses intermediate coordinate systems to avoid directly relating the airplane attitude (the body axes) to the inertial axes. Instead, there is a middle moderating axis system. The presented InvSim algorithm adopts the flight-path angles (Inline graphic) and the wind axes angles (Inline graphic); where it bypasses the singularity problem.

There is another attitude representation that can handle the aforementioned pitch singularity (or gimbal lock), which is using the quaternion notation (rotation quaternions) rather than Euler angles393,394. Mathematically, a quaternion can be viewed as a hyper-complex number or a four-dimensional vector, consisting of one real component and three imaginary components. Thus, a quaternion is an extended version of the classical complex number (having also one real component but only one imaginary component). A unit quaternion is a quaternion that has a norm of exactly 1. Unit quaternions can be used to represent rotations. A unit quaternion may take the following general form395:

graphic file with name 41598_2025_7639_Article_Equ132.gif 109

subject to the normalization condition

graphic file with name 41598_2025_7639_Article_Equ133.gif 110

Because each rotation can be represented by two quaternions (Inline graphic and Inline graphic); at the singularity rotation, the singularity-free quaternion (and related Euler angles solution) can be chosen, while the other is discarded.

Despite the attractive efficiency of the quaternion approach, we prefer the presented algorithm here, which has a number of merits. For example, the presented approach in this study is intuitive and relies totally on real flight quantities. It does not resort to fictitious mathematical quantities and does not require familiarity with detailed quaternion algebra.

Conclusions

In the current study, a numerical algorithm for solving the six-degree-of-freedom inverse simulation (InvSim) flight mechanics problem for an arbitrary fixed-wing airplane was described in detail, starting from the governing equations of motion and elementary geometric relations, and including a submodel for estimating properties of atmospheric air, and steps followed in deriving the underlying system of 35 differential–algebraic equations in 39 variables, with 30 user-defined constant parameters (as well as 8 universal constants).

The algorithm utilizes the four-step Runge–Kutta method, along with second-order finite difference formulas. By specifying the aimed three rectangular coordinates of a trajectory and the aimed profile for the roll angle during a maneuver, the algorithm computes how the four flight control variables (the thrust force; and the three deflection angles of movable control surfaces – rudder, elevators, and ailerons) should change with time to achieve this target maneuver.

The algorithm was demonstrated through a complete test case of a constant-speed two-harmonic double-roll level-flight maneuver for an airplane resembling the Mirage III French fighters. The settings and results of the test problem were discussed, and the effect of the time step size was illustrated.

The work presented here may be extended to more complex operations, for example, through implementing real-time feedback control to compensate for external disturbances (such as gusts), or through accommodating additional geometric features (such as wing flaps).

Supplementary Information

Author contributions

Author Contributions Statement is not applicable for this manuscript. O.M. is the single author of this manuscript.

Data availability

Data generated or analyzed during this study are included in this published article. A supplementary material containing expressions for second time derivatives that were not presented in the main body of the manuscript is provided as a document titled (suppl.pdf).

Declarations

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

The online version contains supplementary material available at 10.1038/s41598-025-07639-6.

References

  • 1.Gamwo, I. K. & Liu, Y. Mathematical modeling and numerical simulation of methane production in a hydrate reservoir. Ind. Eng. Chem. Res.49, 5231–5245. 10.1021/ie901452v (2010). [Google Scholar]
  • 2.Makarovskikh, T. A., Panyukov, A. V. & Savitskiy, E. A. Mathematical models and routing algorithms for CAD technological preparation of cutting processes. Autom. Remote Control78, 868–881. 10.1134/S0005117917050095 (2017). [Google Scholar]
  • 3.Marzouk, O. A. Radiant heat transfer in nitrogen-free combustion environments. Int. J. Nonlinear Sci. Numer. Simul.19, 175–188. 10.1515/ijnsns-2017-0106 (2018). [Google Scholar]
  • 4.Zukowski, M. Mathematical modeling and numerical simulation of a short term thermal energy storage system using phase change material for heating applications. Energy Convers. Manag.48, 155–165. 10.1016/j.enconman.2006.04.017 (2007). [Google Scholar]
  • 5.Jaafar, M. A., Rousse, D. R., Gibout, S. & Bédécarrats, J.-P. A review of dendritic growth during solidification: Mathematical modeling and numerical simulations. Renew. Sustain. Energy Rev.74, 1064–1079. 10.1016/j.rser.2017.02.050 (2017). [Google Scholar]
  • 6.Marzouk, O. A. Temperature-dependent functions of the electron-neutral momentum transfer collision cross sections of selected combustion plasma species. Appl. Sci.13, 11282. 10.3390/app132011282 (2023). [Google Scholar]
  • 7.Parolini, N. & Quarteroni, A. Mathematical models and numerical simulations for the America’s Cup. Comput. Methods Appl. Mech. Eng.194, 1001–1026. 10.1016/j.cma.2004.06.020 (2005). [Google Scholar]
  • 8.Wang, H. G., Yang, W. Q., Senior, P., Raghavan, R. S. & Duncan, S. R. Investigation of batch fluidized-bed drying by mathematical modeling, CFD simulation and ECT measurement. AIChE J.54, 427–444. 10.1002/aic.11406 (2008). [Google Scholar]
  • 9.Marzouk, O. A. Zero carbon ready metrics for a single-family home in the Sultanate of Oman based on EDGE certification system for green buildings. Sustainability15, 13856. 10.3390/su151813856 (2023). [Google Scholar]
  • 10.Yee, H. C. & Sweby, P. K. Nonlinear dynamics and numerical uncertainties in CFD, NASA [United States National Aeronautics and Space Administration]. https://ntrs.nasa.gov/api/citations/19960049992/downloads/19960049992.pdf (Accessed 14 Dec 2024) (ARC [Ames Research Center], 1996).
  • 11.Weiland, C. Computational Space Flight Mechanics10.1007/978-3-642-13583-5 (Accessed 14 Dec 2024) (Springer Science & Business Media, 2010).
  • 12.Yondo, R., Andrés, E. & Valero, E. A review on design of experiments and surrogate models in aircraft real-time and many-query aerodynamic analyses. Prog. Aerosp. Sci.96, 23–61. 10.1016/j.paerosci.2017.11.003 (2018). [Google Scholar]
  • 13.Ghoreyshi, M., Jirásek, A. & Cummings, R. M. Computational investigation into the use of response functions for aerodynamic-load modeling. AIAA J.50, 1314–1327. 10.2514/1.J051428 (2012). [Google Scholar]
  • 14.Wang, Q., Qian, W. & He, K. Unsteady aerodynamic modeling at high angles of attack using support vector machines. Chin. J. Aeronaut.28, 659–668. 10.1016/j.cja.2015.03.010 (2015). [Google Scholar]
  • 15.Marzouk, O. A. Direct numerical simulations of the flow past a cylinder moving with sinusoidal and nonsinusoidal profiles. J. Fluids Eng.131, 121201. 10.1115/1.4000406 (2009). [Google Scholar]
  • 16.Kleinhenz, J. et al. Infrared imaging diagnostics for flame spread over solid surfaces. In 41st Aerospace Sciences Meeting and Exhibit, AIAA 2003-988 10.2514/6.2003-988 (AIAA [American Institute of Aeronautics and Astronautics], 2003).
  • 17.Dogariu, L. E., Dogariu, A., Miles, R. B., Smith, M. S. & Marineau, E. C. Non-intrusive hypersonic freestream and turbulent boundary-layer velocity measurements in AEDC tunnel 9 using FLEET. In 2018 AIAA Aerospace Sciences Meeting, AIAA 2018-1769 10.2514/6.2018-1769 (AIAA [American Institute of Aeronautics and Astronautics], 2018).
  • 18.Chanetz, B. et al. Intrusive measurement techniques. In Experimental Aerodynamics: An Introductory Guide (eds Chanetz, B. et al.) 225–236 10.1007/978-3-030-35562-3_10 (Springer International Publishing, 2020).
  • 19.Marzouk, O. A. & Nayfeh, A. H. Characterization of the flow over a cylinder moving harmonically in the cross-flow direction. Int. J. Non-Linear Mech.45, 821–833. 10.1016/j.ijnonlinmec.2010.06.004 (2010). [Google Scholar]
  • 20.Loh, W. H. T. A second-order theory of entry mechanics into a planetary atmosphere. J. Aerospace Sci.29, 1210–1221. 10.2514/8.9761 (1962). [Google Scholar]
  • 21.Dadallage, N. & Volanka Stepson, W. A. Pre-flight mission simulation model for estimating state-of-energy (SoE) in hybrid aircraft, Master’s Thesis in Systems, Control and Mechatronics, http://hdl.handle.net/20.500.12380/308373 (Accessed 14 Dec 2024) (Chalmers University of Technology, 2024).
  • 22.Marzouk, O. A. Expectations for the role of hydrogen and its derivatives in different sectors through analysis of the four energy scenarios: IEA-STEPS, IEA-NZE, IRENA-PES, and IRENA-1.5°C. Energies17, 646. 10.3390/en17030646 (2024). [Google Scholar]
  • 23.Gordon, R., Ceriotti, M. & Worrall, K. Effects of model fidelity and uncertainty on a model-based attitude controller for satellites with flexible appendages. Acta Astronaut.214, 30–45. 10.1016/j.actaastro.2023.10.011 (2024). [Google Scholar]
  • 24.Gordon, R., Worrall, K. & Ceriotti, M. Broyden’s method for faster inverse simulation. J. Guid. Control. Dyn.47, 1939–1945. 10.2514/1.G007975 (2024). [Google Scholar]
  • 25.Song, J., Wang, Y., Ji, C. & Zhang, H. Real-time optimization control of variable rotor speed based on helicopter/turboshaft engine on-board composite system. Energy301, 131701. 10.1016/j.energy.2024.131701 (2024). [Google Scholar]
  • 26.Marzouk, O. A. Chronologically-ordered quantitative global targets for the energy-emissions-climate nexus, from 2021 to 2050. In 2022 International Conference on Environmental Science and Green Energy (ICESGE) 1–6. 10.1109/ICESGE56040.2022.10180322 (IEEE [Institute of Electrical and Electronics Engineers], 2022).
  • 27.Campa, G., Fravolini, M. L., Napolitano, M. R., Perhinschi, M. G. & Battipede, M. A stochastically optimal feedforward and feedback technique for flight control systems of high performance aircrafts. In Proceedings of the 2004 American Control Conference, vol. 2, 1713–1718 10.23919/ACC.2004.1386826 (2004).
  • 28.Kim, H., Park, Y. M., Aung, P. S., Munir, Md. S. & Hong, C. S. Energy-efficient trajectory and age of information optimization for urban air mobility. In NOMS 2024–2024 IEEE Network Operations and Management Symposium, 1–510.1109/NOMS59830.2024.10575247 (2024).
  • 29.Marzouk, O. A. Portrait of the decarbonization and renewables penetration in Oman’s energy mix, motivated by Oman’s national green hydrogen plan. Energies17, 4769. 10.3390/en17194769 (2024). [Google Scholar]
  • 30.Bridgelall, R. Locating electrified aircraft service to reduce urban congestion. Information15, 186. 10.3390/info15040186 (2024). [Google Scholar]
  • 31.Marzouk, O. A. Compilation of smart cities attributes and quantitative identification of mismatch in rankings. J. Eng.2022, 5981551. 10.1155/2022/5981551 (2022). [Google Scholar]
  • 32.Putra, S. H., Bhandari, R. & Chakraborty, I. Integrated vehicle and subsystems sizing and optimization for advanced air mobility aircraft. In AIAA SCITECH 2024 Forum, AIAA 2024-1085 10.2514/6.2024-1085 (AIAA [American Institute of Aeronautics and Astronautics], 2024).
  • 33.Marzouk, O. A. Coupled differential-algebraic equations framework for modeling six-degree-of-freedom flight dynamics of asymmetric fixed-wing aircraft. Int. J. Appl. Adv. Sci.12, 30–51. 10.21833/ijaas.2025.01.004 (2025). [Google Scholar]
  • 34.Ko, H. Y., Ha, D. J. & Choi, N. Y. Optimal design of the aircraft ABS controller. IFAC Proc. Vol.36, 219–224. 10.1016/S1474-6670(17)35787-7 (2003). [Google Scholar]
  • 35.Couluris, G. J., Signor, D. & Phillips, J. Cruise-efficient short takeoff and landing (CESTOL): potential impact on air traffic operations. https://ntrs.nasa.gov/api/citations/20110008687/downloads/20110008687.pdf (Accessed 14 Dec 2024) (NASA [United States National Aeronautics and Space Administration], ARC [Ames Research Center], 2010).
  • 36.Marzouk, O. A. Land-use competitiveness of photovoltaic and concentrated solar power technologies near the Tropic of Cancer. Sol. Energy243, 103–119. 10.1016/j.solener.2022.07.051 (2022). [Google Scholar]
  • 37.Jeong-Woo Jeon, J.-W. J. et al. Development of a dynamic simulator for braking performance test of aircraft with anti-skid brake system. In: Proceedings of the 2002 IEEE International Symposium on Industrial Electronics, 2002, vol. 2, 518–523 10.1109/ISIE.2002.1026344 (ISIE, 2002).
  • 38.Pan, C.-T. et al. Development of multi-axis crank linkage motion system for synchronized flight simulation with VR immersion. Appl. Sci.11, 3596. 10.3390/app11083596 (2021). [Google Scholar]
  • 39.Wang, X. et al. A fixed-wing UAV formation algorithm based on vector field guidance. IEEE Trans. Autom. Sci. Eng.20, 179–192. 10.1109/TASE.2022.3144672 (2023). [Google Scholar]
  • 40.Stojakovic, P. & Rasuo, B. Single propeller airplane minimal flight speed based upon the lateral maneuver condition. Aerosp. Sci. Technol.49, 239–249. 10.1016/j.ast.2015.12.012 (2016). [Google Scholar]
  • 41.Marzouk, O. A. Lookup tables for power generation performance of photovoltaic systems covering 40 geographic locations (Wilayats) in the Sultanate of Oman, with and without solar tracking, and general perspectives about solar irradiation. Sustainability13, 13209. 10.3390/su132313209 (2021). [Google Scholar]
  • 42.Karelahti, J., Virtanen, K. & Öström, J. Automated generation of realistic near-optimal aircraft trajectories. J. Guid. Control. Dyn.31, 674–688. 10.2514/1.31159 (2008). [Google Scholar]
  • 43.Stojakovic, P. & Rasuo, B. Minimal safe speed of the asymmetrically loaded combat airplane. Aircraft Eng. Aerosp. Technol. Int. J.88, 42–52. 10.1108/AEAT-03-2014-0033 (2016). [Google Scholar]
  • 44.Sri-Jayantha, M. & Stengel, R. F. Determination of nonlinear aerodynamic coefficients using the estimation-before-modeling methods. IFAC Proc. Vol.18, 837–844. 10.1016/S1474-6670(17)60666-9 (1985). [Google Scholar]
  • 45.Marzouk, O. A. & Nayfeh, A. H. Control of ship roll using passive and active anti-roll tanks. Ocean Eng.36, 661–671. 10.1016/j.oceaneng.2009.03.005 (2009). [Google Scholar]
  • 46.Li, H., Zheng, X., He, H. & Liao, L. Design and longitudinal dynamics decoupling control of a tilt-rotor aerial vehicle with high maneuverability and efficiency. IEEE Robot. Autom. Lett.8, 1191–1198. 10.1109/LRA.2022.3231826 (2023). [Google Scholar]
  • 47.Goldman, R. Understanding quaternions. Graph. Models73, 21–49. 10.1016/j.gmod.2010.10.004 (2011). [Google Scholar]
  • 48.Marzouk, O. A. Performance analysis of shell-and-tube dehydrogenation module. Int. J. Energy Res.41, 604–610. 10.1002/er.3637 (2017). [Google Scholar]
  • 49.Zhang, F. Quaternions and matrices of quaternions. Linear Algebra Appl.251, 21–57. 10.1016/0024-3795(95)00543-9 (1997). [Google Scholar]
  • 50.Gade, K. The seven ways to find heading. J. Navig.69, 955–970. 10.1017/S0373463316000096 (2016). [Google Scholar]
  • 51.Espinal, D. et al. Supersonic bi-directional flying wing, Part II: conceptual design of a high speed civil transport. In 48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, AIAA 2010-1393 10.2514/6.2010-1393 (AIAA [American Institute of Aeronautics and Astronautics], 2010).
  • 52.Marzouk, O. A. & Nayfeh, A. H. Loads on a harmonically oscillating cylinder. In ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007), 1755–1774. 10.1115/DETC2007-35562 (ASME [American Society of Mechanical Engineers], 2009).
  • 53.Nolan II, R. C. Wing Rock Prediction Method for a High Performance Fighter Aircraft, Master Thesis https://apps.dtic.mil/sti/pdfs/ADA256438.pdf (Accessed 23 July 2023) (Air Force Institute of Technology, Air University, 1992).
  • 54.Fogarty, L. E. & Howe, R. M. computer mechanization of six-degree of freedom flight equations https://ntrs.nasa.gov/api/citations/19690016854/downloads/19690016854.pdf (Accessed 17 July 2023) (NASA [United States National Aeronautics and Space Administration], 1969).
  • 55.Marzouk, O. A. Tilt sensitivity for a scalable one-hectare photovoltaic power plant composed of parallel racks in Muscat. Cogent. Eng.9, 2029243. 10.1080/23311916.2022.2029243 (2022). [Google Scholar]
  • 56.Kaushik, M. The standard atmosphere. In Theoretical and Experimental Aerodynamics (ed. Kaushik, M.) 3–25 (Springer, 2019). 10.1007/978-981-13-1678-4_1. [Google Scholar]
  • 57.Thuan, D. D. & Son, N. H. Stability of stochastic differential-algebraic equations with delay. Int. J. Syst. Sci.55, 1835–1850. 10.1080/00207721.2024.2321378 (2024). [Google Scholar]
  • 58.Marzouk, O. A. One-way and two-way couplings of CFD and structural models and application to the wake-body interaction. Appl. Math. Model.35, 1036–1053. 10.1016/j.apm.2010.07.049 (2011). [Google Scholar]
  • 59.Huang, Y., Zou, C., Li, Y. & Wik, T. MINN: learning the dynamics of differential-algebraic equations and application to battery modeling. IEEE Trans. Pattern Anal. Mach. Intell.46, 11331–11344. 10.1109/TPAMI.2024.3456475 (2024). [DOI] [PubMed] [Google Scholar]
  • 60.Cartwright, J. H. E. & Piro, O. The dynamics of Runge–Kutta methods. Int. J. Bifurcation Chaos02, 427–449. 10.1142/S0218127492000641 (1992). [Google Scholar]
  • 61.Marzouk, O. A. & Nayfeh, A. H. Reduction of the loads on a cylinder undergoing harmonic in-line motion. Phys. Fluids21, 083103. 10.1063/1.3210774 (2009). [Google Scholar]
  • 62.Butcher, J. C. A history of Runge-Kutta methods. Appl. Numer. Math.20, 247–260. 10.1016/0168-9274(95)00108-5 (1996). [Google Scholar]
  • 63.Bhattacharyya, S. & Thander, A. K. Slab waveguide communication study using Finite Difference Method (FDM) with fourth-order compact scheme. Results Opt.16, 100681. 10.1016/j.rio.2024.100681 (2024). [Google Scholar]
  • 64.Marzouk, O. A. Characteristics of the flow-induced vibration and forces with 1- and 2-DOF vibrations and limiting solid-to-fluid density ratios. J. Vib. Acoust.132, 041013. 10.1115/1.4001503 (2010). [Google Scholar]
  • 65.Dassault Aviation, Our Company Profile │ Dassault Aviation, Dassault Aviation, a Major Player to Aeronautics https://www.dassault-aviation.com/en/group/about-us/company-profile (Accessed 14 Dec 2024) (2024).
  • 66.Szabo, J. South African responses to new Soviet air defence systems in Angola in the 1980s. Scientia Militaria S. Afr. J. Military Stud.52, 123–142 (2024). [Google Scholar]
  • 67.Raol, J. R. & Singh, J. Flight Mechanics Modeling and Analysis 2nd edn. (CRC Press, Taylor & Francis Group, 2023). [Google Scholar]
  • 68.[United States Federal Aviation Administration] FAA, Chapter 3: Basic Flight Maneuvers. In Airplane Flying Handbook (FAA-H-8083-3C), Oklahoma City, Oklahoma, USA, 3.1–3.25 https://www.faa.gov/sites/faa.gov/files/regulations_policies/handbooks_manuals/aviation/airplane_handbook/04_afh_ch3.pdf (Accessed 15 Dec 2024) (2021).
  • 69.Urban, D. et al. Effect of electric ducted fans structural arrangement on their performance characteristics. Appl. Sci.13, 2787. 10.3390/app13052787 (2023). [Google Scholar]
  • 70.Marzouk, O. A. Energy generation intensity (EGI) for parabolic dish/engine concentrated solar power in Muscat, Sultanate of Oman. IOP Conf. Ser. Earth Environ. Sci.1008, 012013. 10.1088/1755-1315/1008/1/012013 (2022). [Google Scholar]
  • 71.Jin, Y., Qian, Y., Zhang, Y. & Zhuge, W. Modeling of ducted-fan and motor in an electric aircraft and a preliminary integrated design. SAE Int. J. Aerosp.11, 115–126. 10.4271/01-11-02-0007 (2018). [Google Scholar]
  • 72.Marzouk, O. A. Urban air mobility and flying cars: Overview, examples, prospects, drawbacks, and solutions. Open Eng.12, 662–679. 10.1515/eng-2022-0379 (2022). [Google Scholar]
  • 73.Vratny, P. C. & Hornung, M. Sizing considerations of an electric ducted fan for hybrid energy aircraft. Transport. Res. Procedia29, 410–426. 10.1016/j.trpro.2018.02.037 (2018). [Google Scholar]
  • 74.Marzouk, O. A. Summary of the 2023 (1st edition) report of TCEP (tracking clean energy progress) by the International Energy Agency (IEA), and proposed process for computing a single aggregate rating. E3S Web Conf.601, 00048. 10.1051/e3sconf/202560100048 (2025). [Google Scholar]
  • 75.Zhang, X., Bowman, C. L., O’Connell, T. C. & Haran, K. S. Large electric machines for aircraft electric propulsion. IET Electr. Power Appl.12, 767–779. 10.1049/iet-epa.2017.0639 (2018). [Google Scholar]
  • 76.Marzouk, O. A. Growth in the worldwide stock of e-mobility vehicles (by technology and by transport mode) and the worldwide stock of hydrogen refueling stations and electric charging points between 2020 and 2022. In Construction Materials and Their Processing, 89–96. 10.4028/p-8IMGm4 (Trans Tech Publications Ltd, 2023).
  • 77.Sahoo, S., Zhao, X. & Kyprianidis, K. A review of concepts, benefits, and challenges for future electrical propulsion-based aircraft. Aerospace7, 44. 10.3390/aerospace7040044 (2020). [Google Scholar]
  • 78.Marzouk, O. A. Recommended LEED-compliant cars, SUVs, vans, pickup trucks, station wagons, and two seaters for smart cities based on the environmental damage index (EDX) and green score. In Innovations in Smart Cities Applications Volume 7 (eds Ben Ahmed, M. et al.), 123–135. 10.1007/978-3-031-53824-7_12 (Springer Nature Switzerland, 2024).
  • 79.Fard, M. T., He, J., Huang, H. & Cao, Y. Aircraft distributed electric propulsion technologies—a review. IEEE Trans. Transport. Electrification8, 4067–4090. 10.1109/TTE.2022.3197332 (2022). [Google Scholar]
  • 80.Marzouk, O. A. Toward more sustainable transportation: Green vehicle metrics for 2023 and 2024 model years. In Intelligent Sustainable Systems, (eds Nagar, A. K. et al.) 261–272. 10.1007/978-981-99-7886-1_23 (Springer Nature Singapore, 2024).
  • 81.Arvidsson, R., Nordelöf, A. & Brynolf, S. Life cycle assessment of a two-seater all-electric aircraft. Int. J. Life Cycle Assess.29, 240–254. 10.1007/s11367-023-02244-z (2024). [Google Scholar]
  • 82.Lowry, J. Future energy without oil and fossil fuel. In Renewable Energy in the Service of Mankind Vol II: Selected Topics from the World Renewable Energy Congress WREC 2014, (ed Sayigh, A) 233–24410.1007/978-3-319-18215-5_21 (Springer International Publishing, 2016).
  • 83.Marzouk, O. A. Jatropha curcas as marginal land development crop in the Sultanate of Oman for producing biodiesel, biogas, biobriquettes, animal feed, and organic fertilizer. Rev. Agric. Sci.8, 109–123. 10.7831/ras.8.0_109 (2020). [Google Scholar]
  • 84.Lowry, J. Air and sea transport without fossil fuels. In Avoiding Carbon Apocalypse Through Alternative Energy: Life After Fossil Fuels (ed Lowry, J.) 107–122 10.1007/978-3-319-52195-4_9 (Springer International Publishing, 2017).
  • 85.Marzouk, O. A. 2030 ambitions for hydrogen, clean hydrogen, and green hydrogen. Eng. Proc.56, 14. 10.3390/ASEC2023-15497 (2023). [Google Scholar]
  • 86.Meunier, F. The greenhouse effect: A new source of energy. Appl. Therm. Eng.27, 658–664. 10.1016/j.applthermaleng.2006.05.028 (2007). [Google Scholar]
  • 87.Marzouk, O. A. Evolution of the (energy and atmosphere) credit category in the LEED green buildings rating system for (building design and construction: new construction), from version 4.0 to version 4.1. J. Infrastruct. Policy Dev.8, 5306. 10.24294/jipd.v8i8.5306 (2024). [Google Scholar]
  • 88.Shinnar, R. & Citro, F. Decarbonization: Achieving near-total energy independence and near-total elimination of greenhouse emissions with available technologies. Technol. Soc.30, 1–16. 10.1016/j.techsoc.2007.10.006 (2008). [Google Scholar]
  • 89.Marzouk, O. A. Thermoelectric generators versus photovoltaic solar panels: Power and cost analysis. Edelweiss Appl. Sci. Technol.8, 406–428. 10.55214/25768484.v8i5.1697 (2024). [Google Scholar]
  • 90.Dornack, A. et al. Optimizing thrust reversal for a small turbojet engine. In 2024 Regional Student Conferences10.2514/6.2024-85769 (AIAA [American Institute of Aeronautics and Astronautics], Multiple Locations, 2024).
  • 91.Bruce, T. & Rytting, T. The analysis of thrust reversal performance. In 19th Joint Propulsion Conference A83–36337 10.2514/6.1983-1316 (AIAA [American Institute of Aeronautics and Astronautics], 1983).
  • 92.Marzouk, O. A. Aerial e-mobility perspective: Anticipated designs and operational capabilities of eVTOL urban air mobility (UAM) aircraft. Edelweiss Appl. Sci. Technol.9, 413–442. 10.55214/25768484.v9i1.4156 (2025). [Google Scholar]
  • 93.Schaefer, J. W., Sagerser, D. R. & Stakolich, E. G. Dynamics of high-bypass-engine thrust reversal using a variable-pitch fan https://ntrs.nasa.gov/api/citations/19770015162/downloads/19770015162.pdf (Accessed 11 Feb 2025) (NASA [United States National Aeronautics and Space Administration], 1977).
  • 94.Wang, X. F., Gao, M. Z. & Wang, H. Z. A uniform rule that can replace the left-hand rule and the right-hand rule in electromagnetics. Appl. Mech. Mater.278–280, 86–89. 10.4028/www.scientific.net/AMM.278-280.86 (2013). [Google Scholar]
  • 95.Juethner, K. et al. Mathuria, mixed whirl modes in numerical rotordynamics analysis. In American Society of Mechanical Engineers Digital Collection10.1115/GT2022-79755 (2022).
  • 96.[Advanced Rotorcraft Technology ART Inc.], FLIGHTLAB - Rotorcraft Simulation Modeling and Analysis Software, https://www.flightlab.com/flightlab.html (Accessed 12 Feb 2025). (2025).
  • 97.Marzouk, O. & Nayfeh, A. Differential/algebraic wake model based on the total fluid force and its direction, and the effect of oblique immersed-body motion on ‘Type-1’ and ‘Type-2’ lock-in. In 47th AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition, AIAA 2009-1112. 10.2514/6.2009-1112 (AIAA [American Institute of Aeronautics and Astronautics], 2009).
  • 98.[Aerospace Control Dynamics LLC] ACD, FLIGHT Dynamics Software, https://acd-eng.com/flight-dynamics (Accessed 12 Feb 2025) (2025).
  • 99.Marzouk, O. A. Facilitating digital analysis and exploration in solar energy science and technology through free computer applications. Eng. Proc.31, 75. 10.3390/ASEC2022-13920 (2022). [Google Scholar]
  • 100.Gear, C. W. differential-algebraic equations. In Computer Aided Analysis and Optimization of Mechanical System Dynamics (ed Haug, E. J.) 323–334 10.1007/978-3-642-52465-3_12 (Springer Berlin Heidelberg, 1984).
  • 101.Marzouk, O. A. Technical review of radiative-property modeling approaches for gray and nongray radiation, and a recommended optimized WSGGM for CO2/H2O-enriched gases. Results Eng.25, 103923. 10.1016/j.rineng.2025.103923 (2025). [Google Scholar]
  • 102.[Parsons Corporation] PC, SatFlyerTM - Complete Flight Dynamic Mission Lifecycle, https://www.parsons.com/products/satflyer (Accessed 12 Feb 2025) (2025).
  • 103.Marzouk, O. A. Levelized cost of green hydrogen (LCOH) in the Sultanate of Oman using H2A-Lite with polymer electrolyte membrane (PEM) electrolyzers powered by solar photovoltaic (PV) electricity. E3S Web Conf.469, 00101. 10.1051/e3sconf/202346900101 (2023). [Google Scholar]
  • 104.Guan, Y. & Verschelde, J. PHClab: A MATLAB/octave interface to PHCpack. In Software for Algebraic Geometry (eds Stillman, M. et al.) 15–32 (Springer, 2008). 10.1007/978-0-387-78133-4_2. [Google Scholar]
  • 105.Marzouk, O. A. Adiabatic flame temperatures for oxy-methane, oxy-hydrogen, air-methane, and air-hydrogen stoichiometric combustion using the NASA CEARUN tool, GRI-Mech 3.0 reaction mechanism, and cantera python package. Eng. Technol. Appl. Sci. Res.13, 11437–11444. 10.48084/etasr.6132 (2023). [Google Scholar]
  • 106.Lessa, I. D. M., Carneiro, G. D. F., Monteiro, M. P. & e Abreu, F. B. A concern visualization approach for improving MATLAB and Octave program comprehension. In 2015 29th Brazilian Symposium on Software Engineering, 130–139 10.1109/SBES.2015.19 (2015).
  • 107.Marzouk, O. A. & Huckaby, E. D. Modeling confined jets with particles and swril. In Machine Learning and Systems Engineering (eds Ao, S.-I. et al.) 243–256 (Springer Netherlands, 2010). 10.1007/978-90-481-9419-3_19. [Google Scholar]
  • 108.Eaton, J. W. GNU Octave and reproducible research. J. Process Control22, 1433–1438. 10.1016/j.jprocont.2012.04.006 (2012). [Google Scholar]
  • 109.Marzouk, O. A. & Nayfeh, A. H. Mitigation of ship motion using passive and active anti-roll tanks. In ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007) 215–229. 10.1115/DETC2007-35571 (ASME [American Society of Mechanical Engineers], 2009).
  • 110.Gilat, A. MATLAB: An Introduction with Applications (Wiley, 2017). [Google Scholar]
  • 111.Eaton, J. W. About GNU Octave. https://octave.org/about (Accessed 15 Dec 2024) (2024).
  • 112.Tsilika, K. A mathematica-based interface for the exploration of inter- and intra-regional financial flows. Mathematics12, 877. 10.3390/math12060877 (2024). [Google Scholar]
  • 113.Savran, M., Sayi, H. & Aydin, L. Mathematica and optimization. In Designing Engineering Structures Using Stochastic Optimization Methods (CRC Press, 2020).
  • 114.Lynch, S. Dynamical Systems with Applications Using Mathematica® (Springer International Publishing, 2017). 10.1007/978-3-319-61485-4. [Google Scholar]
  • 115.Abell, M. L. & Braselton, J. P. Differential Equations with Mathematica (Academic Press, 2022). [Google Scholar]
  • 116.Freeman, J. A. Simulating Neural Networks with Mathematica (Addison-Wesley, 1994). [Google Scholar]
  • 117.Liu, X. & Ma, Y.-Q. AMFlow: A Mathematica package for Feynman integrals computation via auxiliary mass flow. Comput. Phys. Commun.283, 108565. 10.1016/j.cpc.2022.108565 (2023). [Google Scholar]
  • 118.Bera, S. MultiHypExp: A Mathematica package for expanding multivariate hypergeometric functions in terms of multiple polylogarithms. Comput. Phys. Commun.297, 109060. 10.1016/j.cpc.2023.109060 (2024). [Google Scholar]
  • 119.Ruskeepaa, H. Mathematica Navigator: Mathematics (Academic Press, 2009). [Google Scholar]
  • 120.Abell, M. L. & Braselton, J. P. Mathematica by Example (Academic Press, 2017). [Google Scholar]
  • 121.Wolfram, S. The Mathematica Book (Wolfram Research Inc., 2003). [Google Scholar]
  • 122.Adkins, W. A. & Davidson, M. G. Ordinary Differential Equations (Springer New York, 2012). 10.1007/978-1-4614-3618-8. [Google Scholar]
  • 123.Pio, R. Euler angle transformations. IEEE Trans. Autom. Control11, 707–715. 10.1109/TAC.1966.1098430 (1966). [Google Scholar]
  • 124.Shuster, M. D. & Markley, F. L. Generalization of the Euler angles. J. Astronaut. Sci.51, 123–132. 10.1007/BF03546304 (2003). [Google Scholar]
  • 125.Marzouk, O. A. Globalization and diversity requirement in higher education. In The 11th World Multi-Conference on Systemics, Cybernetics and Informatics (WMSCI 2007) - The 13th International Conference on Information Systems Analysis and Synthesis (ISAS 2007) 101–106. 10.14293/PR2199.001469.v1 (IIIS [International Institute of Informatics and Systemics], 2007).
  • 126.Achter, J. L., Halling, M.W. & Womack, K.C. Full-scale reinforced concrete bridge bent condition assessment using forced vibration testing. In: Nondestructive Evaluation of Highways, Utilities, and Pipelines IV 396–403 10.1117/12.387831 (SPIE, 2000).
  • 127.Phlips, P. J., East, R. A. & Pratt, N. H. An unsteady lifting line theory of flapping wings with application to the forward flight of birds. J. Fluid Mech.112, 97–125. 10.1017/S0022112081000311 (1981). [Google Scholar]
  • 128.Marzouk, O. A. & Huckaby, E. D. Simulation of a swirling gas-particle flow using different k-epsilon models and particle-parcel relationships. Eng. Lett.18(1). 10.31219/osf.io/k85nf (2010).
  • 129.Marzouk, O. A. & Huckaby, E. D. Effects of turbulence modeling and parcel approach on dispersed two-phase swirling flow. In World Congress on Engineering and Computer Science 2009 (WCECS 2009) 1–11. https://doi.org/10.48550/arXiv.2501.00037 (IAENG [International Association of Engineers], 2009).
  • 130.Ahmed, M. R. Aerodynamics of a cambered airfoil in ground effect. Int. J. Fluid Mech. Res.32, 157–183. 10.1615/InterJFluidMechRes.v32.i2.30 (2005). [Google Scholar]
  • 131.Triet, N. M., Viet, N. N. & Thang, P. M. Aerodynamic analysis of aircraft wing. VNU J. Sci. Math. Phys.31, 68–75 (2015). [Google Scholar]
  • 132.Marzouk, O. A. & Nayfeh, A. H. New wake models with capability of capturing nonlinear physics. In ASME 2008 27th International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2008) 901–912. 10.1115/OMAE2008-57714 (ASME [American Society of Mechanical Engineers], 2009).
  • 133.Ashraf, M. A., Young, J. & Lai, J. C. S. Reynolds number, thickness and camber effects on flapping airfoil propulsion. J. Fluids Struct.27, 145–160. 10.1016/j.jfluidstructs.2010.11.010 (2011). [Google Scholar]
  • 134.Xu, X. & Lagor, F. D. Quasi-steady effective angle of attack and its use in lift-equivalent motion design. AIAA J.59, 2613–2626. 10.2514/1.J059663 (2021). [Google Scholar]
  • 135.Marzouk, O. A. & Nayfeh, A. H. Hydrodynamic forces on a moving cylinder with time-dependent frequency variations. In 46th AIAA Aerospace Sciences Meeting and Exhibit, AIAA 2008-680. 10.2514/6.2008-680 (AIAA [American Institute of Aeronautics and Astronautics], 2008).
  • 136.Maqsood, A. & Go, T. H. Optimization of hover-to-cruise transition maneuver using variable-incidence wing. J. Aircr.47, 1060–1064. 10.2514/1.44453 (2010). [Google Scholar]
  • 137.Winslow, J., Otsuka, H., Govindarajan, B. & Chopra, I. Basic understanding of airfoil characteristics at low Reynolds numbers (104–105). J. Aircr.55, 1050–1061. 10.2514/1.C034415 (2018). [Google Scholar]
  • 138.Crasta, A. & Khan, S. A. Effect of angle of incidence on stability derivatives of a wing. In Proceedings of the National Conference on Challenges in Research & Technology in the Coming Decades (CRT 2013), Karnataka, India, 3.04 10.1049/cp.2013.2523 (2013).
  • 139.Chandravada, N. Numerical study of flow over NACA 2412 airfoil at various AOA’s. Int. J. Sci. Res. Eng. Trends10, 1028–1033 (2024). [Google Scholar]
  • 140.Karkoulias, D. G., Tzoganis, E. D., Panagiotopoulos, A. G., Acheimastos, S.-G.D. & Margaris, D. P. Computational fluid dynamics study of wing in air flow and air-solid flow using three different meshing techniques and comparison with experimental results in wind tunnel. Computation10, 34. 10.3390/computation10030034 (2022). [Google Scholar]
  • 141.Marzouk, O. A. & Nayfeh, A. H. Simulation, analysis, and explanation of the lift suppression and break of 2:1 force coupling due to in-line structural vibration. In 49th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, AIAA 2008-2309. 10.2514/6.2008-2309 (AIAA [American Institute of Aeronautics and Astronautics], 2008).
  • 142.Ardany, M. D., Pandiangan, P. & Hasan, M. Lift force of airfoil (NACA 0012, NACA 4612, NACA 6612) with variation of angle of attack and camber: computational fluid dynamics study. CERiMRE4, 80. 10.19184/cerimre.v4i2.28372 (2021). [Google Scholar]
  • 143.Heinrich, G.-A., Vogt, S., Lawrance, N. R. J., Stastny, T. J. & Siegwart, R. Y. In-wing pressure measurements for airspeed and airflow angle estimation and high angle-of-attack flight. J. Guid. Control. Dyn.45, 1033–1045. 10.2514/1.G006412 (2022). [Google Scholar]
  • 144.Liu, T., Kuykendoll, K., Rhew, R. & Jones, S. Avian wing geometry and kinematics. AIAA J.44, 954–963. 10.2514/1.16224 (2006). [Google Scholar]
  • 145.Marzouk, O. A. Evolutionary computing applied to design optimization. In ASME 2007 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference (IDETC-CIE 2007), (4–7 September 2007), 995–1003. 10.1115/DETC2007-35502 (ASME [American Society of Mechanical Engineers], 2009).
  • 146.Mineck, R. E. Wind-tunnel investigation of aerodynamic efficiency of three planar elliptical wings with curvature of quarter-chord line, NASA, Scientific and Technical Information Program (1993).
  • 147.Anderson, J. Introduction to Flight 7th edn. (McGraw-Hill Education, 2009). [Google Scholar]
  • 148.Sankaralingam, L. & Ramprasadh, C. A comprehensive survey on the methods of angle of attack measurement and estimation in UAVs. Chin. J. Aeronaut.33, 749–770. 10.1016/j.cja.2019.11.003 (2020). [Google Scholar]
  • 149.Xiao, Q. & Liao, W. Numerical investigation of angle of attack profile on propulsion performance of an oscillating foil. Comput. Fluids39, 1366–1380. 10.1016/j.compfluid.2010.04.006 (2010). [Google Scholar]
  • 150.Tan, P. J. B. & Hsu, M.-H. Diagnosis of faults in wind power generation systems, In 2016 IEEE 11th Conference on Industrial Electronics and Applications (ICIEA), 1459–1462 10.1109/ICIEA.2016.7603815 (2016).
  • 151.Marzouk, O. A. Reduced-order modeling (ROM) of a segmented plug-flow reactor (PFR) for hydrogen separation in integrated gasification combined cycles (IGCC). Processes13, 1455. 10.3390/pr13051455 (2025). [Google Scholar]
  • 152.Figura, L. O. & Teixeira, A. A. Mass density. In Food Physics: Physical Properties - Measurement and Applications (eds Figura, L. O. & Teixeira, A. A.) 59–100 (Springer International Publishing, 2023). 10.1007/978-3-031-27398-8_2. [Google Scholar]
  • 153.Chassaing, P. One-dimensional steady flows of an ideal compressible gas In Fundamentals of Fluid Mechanics: For Scientists and Engineers (ed Chassaing, P.), 315–354 10.1007/978-3-031-10086-4_8 (Springer International Publishing, 2022).
  • 154.Marzouk, O. A. Dataset of total emissivity for CO2, H2O, and H2O-CO2 mixtures; over a temperature range of 300–2900 K and a pressure-pathlength range of 001-50 atm.m. Data Brief59, 111428. 10.1016/j.dib.2025.111428 (2025). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Duan, K., Dai, Y., Wu, J. & Zhang, H. A pressure-temperature coupled probe method for pneumatic measurement in non-ideal compressible gas flows. Measurement246, 116571. 10.1016/j.measurement.2024.116571 (2025). [Google Scholar]
  • 156.Marzouk, O. A. Power density and thermochemical properties of hydrogen magnetohydrodynamic (H2MHD) generators at different pressures, seed types, seed levels, and oxidizers. Hydrogen6, 31. 10.3390/hydrogen6020031 (2025). [Google Scholar]
  • 157.Cornelius, K. C. & Srinivas, K. Isentropic compressible flow for non-ideal gas models for a venturi. J. Fluids Eng.126, 238–244. 10.1115/1.1677499 (2004). [Google Scholar]
  • 158.Neely, R. H., Bollech, T. V., Westrick, G. C. & Graham, R. R. Experimental and calculated characteristics of several naca 44-series wings with aspect ratios of 8, 10, and 12 and taper ratios of 2.5 and 3.5, https://ntrs.nasa.gov/api/citations/19930081871/downloads/19930081871.pdf (Accessed 22 Feb 2025) (NACA [United States National Advisory Committee for Aeronautics], 1947).
  • 159.Marzouk, O. A. Wind speed Weibull model identification in Oman, and computed normalized annual energy production (NAEP) from wind turbines based on data from weather stations. Eng. Rep.7, e70089. 10.1002/eng2.70089 (2025). [Google Scholar]
  • 160.Goodson, K. W., Morrison Jr., W. D., Aerodynamic characteristics of a wing with unswept quarter-chord line, aspect ratio 4, taper ratio 0.6, and NACA 65A006 airfoil section, https://ntrs.nasa.gov/api/citations/19930085977/downloads/19930085977.pdf (Accessed 11 Feb 2025). (NACA [United States National Advisory Committee for Aeronautics], Langley Air Force Base, 1949).
  • 161.Wang, J. J., Li, Y. C. & Choi, K.-S. Gurney flap—Lift enhancement, mechanisms and applications. Prog. Aerosp. Sci.44, 22–47. 10.1016/j.paerosci.2007.10.001 (2008). [Google Scholar]
  • 162.Rodriguez, D. L., Aftosmis, M. J., Nemec, M. & Anderson, G. R. Optimization of flexible wings with distributed flaps at off-design conditions. J. Aircr.53, 1731–1745. 10.2514/1.C033535 (2016). [Google Scholar]
  • 163.Behal, A., Rao, V. M., Marzocca, P. & Kamaludeen, M. Adaptive control for a nonlinear wing section with multiple flaps. J. Guid. Control. Dyn.29, 744–749. 10.2514/1.18182 (2006). [Google Scholar]
  • 164.Qin, Y., Liu, P., Qu, Q. & Guo, H. Numerical study of aerodynamic forces and flow physics of a delta wing in dynamic ground effect. Aerosp. Sci. Technol.51, 203–221. 10.1016/j.ast.2016.02.007 (2016). [Google Scholar]
  • 165.Secco, N. R. & Mattos, B. S. D. Artificial neural networks to predict aerodynamic coefficients of transport airplanes. AEAT89, 211–230. 10.1108/AEAT-05-2014-0069 (2017). [Google Scholar]
  • 166.Sogukpinar, H. & Bozkurt, I. Calculation of aerodynamic performance characteristics of airplane wing and comparing with the experimental measurement. Int. J. Eng. Technol. IJET1, 83–87. 10.19072/ijet.105718 (2015). [Google Scholar]
  • 167.Melnikov, V. G. Inertia tensors and centres of masses identification at semiprogram precession motions. In 2012 IEEE International Conference on Control Applications, 494–497 10.1109/CCA.2012.6402471 (IEEE, 2012).
  • 168.Asai, T., Araki, Y. & Ikago, K. Energy harvesting potential of tuned inertial mass electromagnetic transducers. Mech. Syst. Signal Process.84, 659–672. 10.1016/j.ymssp.2016.07.048 (2017). [Google Scholar]
  • 169.Pintão, C. A. F. Measurement of the rotational inertia of bodies by using mechanical spectroscopy. J. Mech. Sci. Technol.28, 4011–4020. 10.1007/s12206-014-0914-8 (2014). [Google Scholar]
  • 170.Yiu, Y., Weston, E. & Loh, L. Mode reduction of flexible appendages via transfer functions and correction vectors. In 32nd Structures, Structural Dynamics, and Materials Conference, AIAA-91-1182-CP 10.2514/6.1991-1182 (AIAA [American Institute of Aeronautics and Astronautics], 1991).
  • 171.Baker, D. W. & Haynes, W. Statics: Products of Inertia, Baker and Haynes, Online https://engineeringstatics.org/products-of-inertia.html (Accessed 19 Dec 2024) (2024).
  • 172.Rucker, C. & Wensing, P. M. Smooth parameterization of rigid-body inertia. IEEE Robot. Autom. Lett.7, 2771–2778. 10.1109/LRA.2022.3144517 (2022). [Google Scholar]
  • 173.Atchonouglo, K., Dumitriu, D., Monnet, T. & Vallée, C. Matrix formulation of the motion equations of a rigid body and identification of the ten inertia characteristics. Proc. Appl. Math. Mech.8, 10109–10110. 10.1002/pamm.200810109 (2008). [Google Scholar]
  • 174.Beatty, M. F. The moment of inertia tensor. In Principles of Engineering Mechanics, 355–404 10.1007/978-0-387-31255-2_5 (Springer US, 2006).
  • 175.Lorenzetti, J. S., Bañuelos, L., Clarke, R., Murillo, O. J. & Bowers, A. Determining products of inertia for small scale UAVs. In 55th AIAA Aerospace Sciences Meeting10.2514/6.2017-0547 (American Institute of Aeronautics and Astronautics, 2017).
  • 176.Turner, H. L. Measurement of the moments of inertia of an airplane by a simplified method. https://ntrs.nasa.gov/api/citations/19930082849/downloads/19930082849.pdf (Accessed 26 Aug 2024) (NACA [United States National Advisory Committee for Aeronautics], 1950).
  • 177.Braca, P., Willett, P., LePage, K., Marano, S. & Matta, V. Bayesian tracking in underwater wireless sensor networks with port-starboard ambiguity. IEEE Trans. Signal Process.62, 1864–1878. 10.1109/TSP.2014.2305640 (2014). [Google Scholar]
  • 178.Hull, D. G. Fundamentals of Airplane Flight Mechanics (Springer, 2007). 10.1007/978-3-540-46573-7. [Google Scholar]
  • 179.Meier, H. A. & Gale, P. A. Alternatives in aircraft carrier design. Nav. Eng. J.89, 19–30. 10.1111/j.1559-3584.1977.tb03846.x (1977). [Google Scholar]
  • 180.Tewari, A. Flight of airplanes and gliders: vertical plane. In Basic Flight Mechanics, 43–72. 10.1007/978-3-319-30022-1_3 (Springer International Publishing, 2016).
  • 181.Svozil, K. Early UFO sagas and legends prior to trinity (July 1945). In UFOs: Unidentified Aerial Phenomena: Observations, Explanations and Speculations (ed. Svozil, K.) 3–13 10.1007/978-3-031-34398-8_1 (Springer International Publishing, 2023).
  • 182.Beknalkar, S., Bryant, M. & Mazzoleni, A. Algorithm for locomotion mode selection, energy estimation and path planning for a multi-terrain screw-propelled vehicle for arctic exploration. In 2024 IEEE International Conference on Advanced Intelligent Mechatronics (AIM), 1462–1467 10.1109/AIM55361.2024.10636948 (2024).
  • 183.Rizzi, S. A., Letica, S. J., Boyd, D. D. & Lopes, L. V. Prediction of noise-power-distance data for urban air mobility vehicles. J. Aircr.61, 166–182. 10.2514/1.C037435 (2024). [Google Scholar]
  • 184.Kim, J.-H. et al. Establishment of a pressure variation model for the state estimation of an underwater vehicle. Appl. Sci.14, 970. 10.3390/app14030970 (2024). [Google Scholar]
  • 185.Slater, T. F. The promises of studying cultural archeoastronomy across the Pacific Basin as a new scholarly science education research agenda. STEM Educ. Rev.10.54844/stemer.2023.0478 (2024). [Google Scholar]
  • 186.Phillips, W., Hailey, C. & Gebert, G. A review of attitude kinematics for aircraft flight simulation. In Modeling and Simulation Technologies Conference, AIAA-2000-4302 10.2514/6.2000-4302 (AIAA [American Institute of Aeronautics and Astronautics], 2000).
  • 187.Kato, O. Some basic considerations on angles describing airplane flight maneuvers. J. Guid. Control. Dyn.17, 378–384. 10.2514/3.21208 (1994). [Google Scholar]
  • 188.Bugajski, D. J. & Enns, D. F. Nonlinear control law with application to high angle-of-attack flight. J. Guid. Control. Dyn.15, 761–767. 10.2514/3.20902 (1992). [Google Scholar]
  • 189.Johansen, T. A., Cristofaro, A., Sørensen, K., Hansen, J. M. & Fossen, T. I. On estimation of wind velocity, angle-of-attack and sideslip angle of small UAVs using standard sensors. In 2015 International Conference on Unmanned Aircraft Systems (ICUAS), 510–519. 10.1109/ICUAS.2015.7152330 (2015).
  • 190.Popowski, S. & Dąbrowski, W. Measurement and estimation of the angle of attack and the angle of sideslip. Aviation19, 19–24. 10.3846/16487788.2015.1015293 (2015). [Google Scholar]
  • 191.Valasek, J. et al. Derived angle of attack and sideslip angle characterization for general aviation. J. Guid. Control. Dyn.43, 1039–1055. 10.2514/1.G004010 (2020). [Google Scholar]
  • 192.Klein, V. & Noderer, K. D. Modeling of aircraft unsteady aerodynamic characteristics. Part 2: Parameters estimated from wind tunnel data. https://ntrs.nasa.gov/api/citations/19950021418/downloads/19950021418.pdf (Accessed 11 Feb 2025) (NASA [United States National Aeronautics and Space Administration], 1995).
  • 193.Grauer, J. A. & Morelli, E. A. A generic nonlinear aerodynamic model for aircraft. In AIAA Atmospheric Flight Mechanics Conference, AIAA 2014-0542. 10.2514/6.2014-0542 (AIAA [American Institute of Aeronautics and Astronautics], 2014).
  • 194.Schmidt, L. V. Introduction to Aircraft Flight Dynamics, 1st ed https://books.google.com.om/books?id=AUWCXVepMKEC (American Institute of Aeronautics and Astronautics, 1998).
  • 195.Meku, A. A., Nageswara Rao, D. K., Mebratu, M. M. & Getnet, L. A. Evaluation of the impact of wing span and wing chord length on the aerodynamic performance of Cessna 172-R aircraft. In Advancement of Science and Technology, (eds Woldegiorgis, B. H. et al.) 419–434 10.1007/978-3-031-33610-2_24 (Springer Nature Switzerland, 2023).
  • 196.Spedding, G. R. & McArthur, J. Span efficiencies of wings at low Reynolds numbers. J. Aircr.47, 120–128. 10.2514/1.44247 (2010). [Google Scholar]
  • 197.Marzouk, O. A. & Nayfeh, A. H. Fluid forces and structure-induced damping of obliquely-oscillating offshore structures. In The Eighteenth (2008) International Offshore and Polar Engineering Conference (ISOPE-2008), 460–468. 10.31219/osf.io/rzkgm_v1 (ISOPE [International Society of Offshore and Polar Engineers], 2008).
  • 198.Yates, A. H. Notes on the mean aerodynamic chord and the mean aerodynamic centre of a wing. J. R. Aeronaut. Soc.56, 461–474. 10.1017/S0368393100129311 (1952). [Google Scholar]
  • 199.Vogeltanz, T. Application for calculation of mean aerodynamic chord of arbitrary wing planform. AIP Conf. Proc.1738, 120018. 10.1063/1.4951901 (2016). [Google Scholar]
  • 200.Shmelova, T. et al. (eds) Automated Systems in the Aviation and Aerospace Industries (IGI Global, 2019). 10.4018/978-1-5225-7709-6. [Google Scholar]
  • 201.Houghton, E. L., Carpenter, P. W., Collicott, S. H. & Valentine, D. T. Chapter 1 - basic concepts and definitions. In Aerodynamics for Engineering Students (Seventh Edition) (eds Carpenter, W. et al.) 1–86 10.1016/B978-0-08-100194-3.00001-8 (Butterworth-Heinemann, 2017).
  • 202.Torenbeek, E. Preliminary tailplane design. In Synthesis of Subsonic Airplane Design: An Introduction to the Preliminary Design of Subsonic General Aviation and Transport Aircraft, with Emphasis on Layout, Aerodynamic Design, Propulsion and Performance (ed. Torenbeek, E.) 303–339. 10.1007/978-94-017-3202-4_9 (Springer Netherlands, 1982).
  • 203.Cook, M. V. Chapter 2 - Systems of Axes and Notation. In Flight Dynamics Principles (Third Edition), (ed. Cook, M. V.) 13–32 10.1016/B978-0-08-098242-7.00002-X (Butterworth-Heinemann, 2013).
  • 204.Custodio, D., Henoch, C. W. & Johari, H. Aerodynamic characteristics of finite span wings with leading-edge protuberances. AIAA J.53, 1878–1893. 10.2514/1.J053568 (2015). [Google Scholar]
  • 205.Gallman, J. W., Smith, S. C. & Kroo, I. M. Optimization of joined-wing aircraft. J. Aircr.30, 897–905. 10.2514/3.46432 (1993). [Google Scholar]
  • 206.Mendelsohn, J. M., Kemp, A. C., Biggs, H. C., Biggs, R. & Brown, C. J. Wing areas, wing loadings and wing spans of 66 species of African raptors. Ostrich60, 35–42. 10.1080/00306525.1989.9634503 (1989). [Google Scholar]
  • 207.Santos, P., Sousa, J. & Gamboa, P. Variable-span wing development for improved flight performance. J. Intell. Mater. Syst. Struct.28, 961–978. 10.1177/1045389X15595719 (2017). [Google Scholar]
  • 208.Tucker, V. A. Gliding birds: the effect of variable wing span. J. Exp. Biol.133, 33–58. 10.1242/jeb.133.1.33 (1987). [Google Scholar]
  • 209.Aslanov, V. S. Removal of large space debris by a tether tow. In Rigid Body Dynamics for Space Applications, 255–356. 10.1016/B978-0-12-811094-2.00005-4 (Elsevier, 2017).
  • 210.Mialon, B. et al. Validation of numerical prediction of dynamic derivatives: The DLR-F12 and the Transcruiser test cases. Prog. Aerosp. Sci.47, 674–694. 10.1016/j.paerosci.2011.08.010 (2011). [Google Scholar]
  • 211.Marzouk, O. A. Flow control using bifrequency motion. Theoret. Comput. Fluid Dyn.25, 381–405. 10.1007/s00162-010-0206-6 (2011). [Google Scholar]
  • 212.Islas-Narvaez, E. A., Ituna-Yudonago, J. F., Ramos-Velasco, L. E., Vega-Navarrete, M. A. & Garcia-Salazar, O. Design and determination of aerodynamic coefficients of a tail-sitter aircraft by means of CFD numerical simulation. Machines11, 17. 10.3390/machines11010017 (2022). [Google Scholar]
  • 213.Jung, B. K. & Rezgui, D. Sectional leading edge vortex lift and drag coefficients of autorotating samaras. Aerospace10, 414. 10.3390/aerospace10050414 (2023). [Google Scholar]
  • 214.Zheng, C., Jin, Z., Dong, Q. & Yang, Z. Numerical investigation of the influences of ridge ice parameters on lift and drag coefficients of airfoils through design of experiments. Adv. Mech. Eng.16, 16878132231226056. 10.1177/16878132231226056 (2024). [Google Scholar]
  • 215.Marzouk, O. A. Contrasting the Cartesian and polar forms of the shedding-induced force vector in response to 12 subharmonic and superharmonic mechanical excitations. Fluid Dyn. Res.42, 035507. 10.1088/0169-5983/42/3/035507 (2010). [Google Scholar]
  • 216.Asadi Asrami, E. & Moonesun, M. Numerical and experimental investigation of the hydrodynamic Lift and Drag coefficients of a solar-powered AUV in near-surface mode. Int. J. Maritime Technol.18, 1–14. 10.61186/ijmt.18.1 (2023). [Google Scholar]
  • 217.Batina, J. T. Effects of airfoil shape, thickness, camber, and angle of attack on calculated transonic unsteady airloads. https://ntrs.nasa.gov/api/citations/19850010648/downloads/19850010648.pdf (Accessed 27 Aug 2024) (NASA [United States National Aeronautics and Space Administration], 1985).
  • 218.Anggraeni, R. Effect of angle of attack on pressure and lift coefficient of ONERA OA206 wing model using computational fluid dynamics method. CERiMRE2, 81. 10.19184/cerimre.v2i2.27374 (2019). [Google Scholar]
  • 219.Phillips, W. F. Lifting-line analysis for twisted wings and washout-optimized wings. J. Aircr.41, 128–136. 10.2514/1.262 (2004). [Google Scholar]
  • 220.Qu, Q., Wang, W., Liu, P. & Agarwal, R. K. Airfoil aerodynamics in ground effect for wide range of angles of attack. AIAA J.53, 1048–1061. 10.2514/1.J053366 (2015). [Google Scholar]
  • 221.Halfman, R. L., Johnson, H. C. & Haley, S. M. Evaluation of high-angle-of-attack aerodynamic-derivative data and stall-flutter prediction techniques (https://ntrs.nasa.gov/api/citations/19930083408/downloads/19930083408.pdf (Accessed 25 Aug 2024) (NACA [United States National Advisory Committee for Aeronautics], 1951).
  • 222.Ham, N. D. & Garelick, M. S. Dynamic stall considerations in helicopter rotors. J. Am. Helicopter Soc.13, 49–55. 10.4050/JAHS.13.49 (1968). [Google Scholar]
  • 223.Jahnke, C. C. & Culick, F. E. C. Application of bifurcation theory to the high-angle-of-attack dynamics of the F-14. J. Aircr.31, 26–34. 10.2514/3.46451 (1994). [Google Scholar]
  • 224.Petrilli, J. L., Paul, R. C., Gopalarathnam, A. & Frink, N. T. A CFD database for airfoils and wings at post-stall angles of attack. In 31st AIAA Applied Aerodynamics Conference10.2514/6.2013-2916 (American Institute of Aeronautics and Astronautics, 2013).
  • 225.Sosa, R., Artana, G., Moreau, E. & Touchard, G. Stall control at high angle of attack with plasma sheet actuators. Exp. Fluids42, 143–167. 10.1007/s00348-006-0227-5 (2006). [Google Scholar]
  • 226.Abbott, I. H. & Von Doenhoff, A. E. Theory of Wing Sections: Including a Summary of Airfoil Datahttps://store.doverpublications.com/products/9780486605869 (Accessed 18 Aug 2024) (Dover Publications, 1959).
  • 227.Andreu Angulo, I. & Ansell, P. J. Influence of aspect ratio on dynamic stall of a finite wing. AIAA J.57, 2722–2733. 10.2514/1.J057792 (2019). [Google Scholar]
  • 228.Hantrais-Gervois, J.-L. & Destarac, D. Drag polar invariance with flexibility. J. Aircr.52, 997–1001. 10.2514/1.C033193 (2015). [Google Scholar]
  • 229.Mavris, D. N. & Qiu, S. An Improved Process for the Generation of Drag Polars for use in Conceptual/Preliminary Design, 1999-01–5641 10.4271/1999-01-5641 (1999).
  • 230.Sun, J., Hoekstra, J. M. & Ellerbroek, J. Estimating aircraft drag polar using open flight surveillance data and a stochastic total energy model. Transport. Res. Part C Emerg. Technol.114, 391–404. 10.1016/j.trc.2020.01.026 (2020). [Google Scholar]
  • 231.McClamroch, N. H. Steady Aircraft Flight and Performance, 1st ed. https://books.google.com.om/books?id=DKK7m8o7_ZkC (Princeton University Press, 2011).
  • 232.Daidzic, N. Efficient general computational method for estimation of standard atmosphere parameters. IJAAA10.15394/ijaaa.2015.1053 (2014). [Google Scholar]
  • 233.Kuprikov, N. M. International standard atmosphere—A tool for technological measurement sovereignty in the aerospace industry. E3S Web Conf.460, 7022. 10.1051/e3sconf/202346007022 (2023). [Google Scholar]
  • 234.Mikhailov, A. A., Tidblad, J. & Kucera, V. The classification system of ISO 9223 standard and the dose-response functions assessing the corrosivity of outdoor atmospheres. Prot. Met.40, 541–550. 10.1023/B:PROM.0000049517.14101.68 (2004). [Google Scholar]
  • 235.Minzner, R. A. The 1976 standard atmosphere and its relationship to earlier standards. Rev. Geophys.15, 375–384. 10.1029/RG015i003p00375 (1977). [Google Scholar]
  • 236.Santana, J. J. et al. Shortcomings of International standard ISO 9223 for the classification, determination, and estimation of atmosphere corrosivities in subtropical archipelagic conditions—the case of the Canary Islands (Spain). Metals9, 1105. 10.3390/met9101105 (2019). [Google Scholar]
  • 237.[International Civil Aviation Organization] ICAO, [Langley Aeronautical Laboratory] LAL, Standard Atmosphere - Tables and Data for Altitudes to 65,800 Feet, https://ntrs.nasa.gov/api/citations/19930090991/downloads/19930090991.pdf (Accessed 20 Mar 2023) (NACA [United States National Advisory Committee for Aeronautics], 1955).
  • 238.Meriwether, J. W. & Gerrard, A. J. Mesosphere inversion layers and stratosphere temperature enhancements. Rev. Geophys.10.1029/2003RG000133 (2004). [Google Scholar]
  • 239.Scherllin-Pirscher, B. et al. The vertical and spatial structure of ENSO in the upper troposphere and lower stratosphere from GPS radio occultation measurements. Geophys. Res. Lett.10.1029/2012GL053071 (2012). [Google Scholar]
  • 240.Gorbushin, A. R. Air humidity effect on flow parameters in subsonic and transonic wind tunnels. TsAGI Sci. J.49, 1–12. 10.1615/TsAGISciJ.2018026783 (2018). [Google Scholar]
  • 241.Woody, A. I. How is the ideal gas law explanatory?. Sci. Educ.22, 1563–1580. 10.1007/s11191-011-9424-6 (2013). [Google Scholar]
  • 242.Marzouk, O. A. Estimated electric conductivities of thermal plasma for air-fuel combustion and oxy-fuel combustion with potassium or cesium seeding. Heliyon10, e31697. 10.1016/j.heliyon.2024.e31697 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Struchtrup, H. Thermodynamics and Energy Conversion, 1st ed. 10.1007/978-3-662-43715-5 (Accessed 21 Aug 2021) (Springer, 2014).
  • 244.Hasan, W. S. A., Hassan, A. S. M. & Shukri, M. A. Assessment of wind power potential and economic viability at Al-Hodeidah in Yemen: Supplying local communities with electricity using wind energy. Energy Rep.12, 2981–2996. 10.1016/j.egyr.2024.08.073 (2024). [Google Scholar]
  • 245.Chadee, X. T. & Clarke, R. M. Air density climate of two caribbean tropical islands and relevance to wind power. Int. Sch. Res. Not.2013, 378027. 10.1155/2013/378027 (2013). [Google Scholar]
  • 246.Brutsaert, W. On a derivable formula for long-wave radiation from clear skies. Water Resour. Res.11, 742–744. 10.1029/WR011i005p00742 (1975). [Google Scholar]
  • 247.Zahariea, D. & Husaru, D. E. Atmospheric air density analysis with Meteo-40S wind monitoring system. MATEC Web Conf.112, 07020. 10.1051/matecconf/201711207020 (2017). [Google Scholar]
  • 248.Sanno, K. & Rao, K. V. S. Effect of air density on kite power. In 2015 International Conference on Technological Advancements in Power and Energy (TAP Energy), Kollam, India, 132–137 10.1109/TAPENERGY.2015.7229605 (2015).
  • 249.Torenbeek, E. Advanced Aircraft Design: Conceptual Design, Analysis and Optimization of Subsonic Civil Airplanes 1st edn. (Wiley, 2013). 10.1002/9781118568101. [Google Scholar]
  • 250.Marzouk, O. A. Assessment of global warming in Al Buraimi, sultanate of Oman based on statistical analysis of NASA POWER data over 39 years, and testing the reliability of NASA POWER against meteorological measurements. Heliyon7, e06625. 10.1016/j.heliyon.2021.e06625 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Yuksel, O. & Aygün, H. Comparative performance analysis of a turboprop engine used in regional aircraft by considering design and flight conditions. Aircr. Eng. Aerosp. Technol.10.1108/AEAT-07-2024-0194 (2025). [Google Scholar]
  • 252.Marzouk, O. A. Condenser pressure influence on ideal steam rankine power vapor cycle using the python extension package cantera for thermodynamics. Eng. Technol. Appl. Sci. Res.14, 14069–14078. 10.48084/etasr.7277 (2024). [Google Scholar]
  • 253.Oksuztepe, E., Bayrak, Z. U. & Kaya, U. Effect of flight level to maximum power utilization for PEMFC/supercapacitor hybrid UAV with switched reluctance motor thruster. Int. J. Hydrogen Energy48, 11003–11016. 10.1016/j.ijhydene.2022.12.160 (2023). [Google Scholar]
  • 254.Matsui, M., Michishita, K. & Yokoyama, S. Influence of the −10 °C isotherm altitudes on winter lightning incidence at wind turbines in coastal areas of the Sea of Japan. Atmos. Res.296, 107071. 10.1016/j.atmosres.2023.107071 (2023). [Google Scholar]
  • 255.Tamesue, K. et al. A machine learning-based non-precipitating clouds estimation for THz dual-frequency radar. In 2023 IEEE Conference on Antenna Measurements and Applications (CAMA), 376–380 10.1109/CAMA57522.2023.10352672 (2023).
  • 256.Levchenko, M. M. & Levchenko, M. A. Information technology procedures of error identification in metrological documents specifying the standard atmosphere parameters in terms of aircraft. Russ. Aeronaut.66, 182–192. 10.3103/S1068799823010257 (2023). [Google Scholar]
  • 257.Ferreira, A. P. & Gimeno, L. Determining precipitable water vapour from upper-air temperature, pressure and geopotential height. Q. J. R. Meteorol. Soc.150, 484–522. 10.1002/qj.4609 (2024). [Google Scholar]
  • 258.Kitajima, T., Akiyama, K., Suzuki, H. & Yasuno, T. Cloud distribution forecasting model using ground altitude information and CNN. In Progressive and Integrative Ideas and Applications of Engineering Systems Under the Framework of IOT and AI (ed. Ma, Y.), 134–145 10.1007/978-981-99-6303-4_11 (Springer Nature Singapore, 2024).
  • 259.Simonetti, M. & Crespillo, O. G. Geodetic altitude from barometer and weather data for GNSS integrity monitoring in aviation. NAVIGATION J. Inst. Navig.10.33012/navi.637 (2024). [Google Scholar]
  • 260.Stober, G., Baumgarten, K., McCormack, J. P., Brown, P. & Czarnecki, J. Comparative study between ground-based observations and NAVGEM-HA analysis data in the mesosphere and lower thermosphere region. Atmos. Chem. Phys.20, 11979–12010. 10.5194/acp-20-11979-2020 (2020). [Google Scholar]
  • 261.Matyja, T., Stanik, Z. & Kubik, A. Automatic correction of barometric altimeters using additional air temperature and humidity measurements. GPS Solut28, 40. 10.1007/s10291-023-01582-7 (2023). [Google Scholar]
  • 262.Acar, Z. & Eşsiz, M. Synoptic analysis of the January 2004 snowstorm: Example of Çanakkale. IJEG.8, 11–18. 10.26833/ijeg.988115 (2023). [Google Scholar]
  • 263.Atasay, N., Atmanli, A. & Yilmaz, N. Liquid cooling flow field design and thermal analysis of proton exchange membrane fuel cells for space applications. Int. J. Energy Res.2023, 7533993. 10.1155/2023/7533993 (2023). [Google Scholar]
  • 264.Matsui, M., Honjo, N., Michishita, K. & Yokoyama, S. Characteristics of lightning discharges to wind turbines and weather conditions at upper air in winter in Japan. In 2023 12th Asia-Pacific International Conference on Lightning (APL), 1–6 10.1109/APL57308.2023.10181380 (2023).
  • 265.Yuan, P. et al. Chapter 10 - GNSS application for weather and climate change monitoring. In GNSS Monitoring of the Terrestrial Environment, (eds Aoki, Y. & Kreemer, C.), 189–204 10.1016/B978-0-323-95507-2.00006-2 (Elsevier, 2024).
  • 266.Marzouk, O. A. Energy generation intensity (EGI) of solar updraft tower (SUT) power plants relative to CSP plants and PV power plants using the new energy simulator “Aladdin”. Energies17, 405. 10.3390/en17020405 (2024). [Google Scholar]
  • 267.Daidzic, N. Long and short-range air navigation on spherical Earth. IJAAA10.15394/ijaaa.2017.1160 (2017). [Google Scholar]
  • 268.[United States National Aeronautics and Space Administration] NASA, Earth Fact Sheet. https://nssdc.gsfc.nasa.gov/planetary/factsheet/earthfact.html (Accessed 22 Aug 2024) (2024).
  • 269.Orlov, S. Genesis of the planet earth. Int. J. Sci. Res. Civil Eng.2, 51–59 (2018). [Google Scholar]
  • 270.Mahony, J. D. Gauging the earth. Math. Gaz.97, 413–420. 10.1017/S0025557200000140 (2013). [Google Scholar]
  • 271.Deng, Y., Ridley, A. J. & Wang, W. Effect of the altitudinal variation of the gravitational acceleration on the thermosphere simulation. J. Geophys. Res.113, 2008JA013081. 10.1029/2008JA013081 (2008). [Google Scholar]
  • 272.Prasad, S., Ivanco, M. L., Ivanco, T. & Ancel, E. Characterization of atmospheric turbulence as a function of altitude. In 9th AIAA Atmospheric and Space Environments Conference10.2514/6.2017-3064 (American Institute of Aeronautics and Astronautics, 2017).
  • 273.Samuels, M. P. The effects of flight and altitude. Arch. Dis. Child.89, 448–455. 10.1136/adc.2003.031708 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Sforza, P. Chapter 3 - Fuselage design. In Commercial Airplane Design Principles (ed Sforza, P.) 47–79 10.1016/B978-0-12-419953-8.00003-6 (Butterworth-Heinemann, 2014).
  • 275.AN Aviation, Cruising altitude for aircraft—challenges and considerations, https://an.aero/understanding-the-significance-of-cruising-altitude (Accessed 1 Oct 2024) (AN Aviation Services Co., 2024).
  • 276.Arnold, F., Scheid, J., Stilp, Th., Schlager, H. & Reinhardt, M. E. Measurements of jet aircraft emissions at cruise altitude I: The odd-nitrogen gases NO, NO2, HNO2 and HNO3. Geophys. Res. Lett.19, 2421–2424. 10.1029/92GL02926 (1992). [Google Scholar]
  • 277.Schulte, P. et al. NOx emission indices of subsonic long-range jet aircraft at cruise altitude: In situ measurements and predictions. J. Geophys. Res. Atmos.102, 21431–21442. 10.1029/97JD01526 (1997). [Google Scholar]
  • 278.Ng, H. K., Sridhar, B. & Grabbe, S. Optimizing aircraft trajectories with multiple cruise altitudes in the presence of winds. J. Aerosp. Inf. Syst.11, 35–47. 10.2514/1.I010084 (2014). [Google Scholar]
  • 279.Petzold, A., Döpelheuer, A., Brock, C. A. & Schröder, F. In situ observations and model calculations of black carbon emission by aircraft at cruise altitude. J. Geophys. Res. Atmos.104, 22171–22181. 10.1029/1999JD900460 (1999). [Google Scholar]
  • 280.Lee, S. H., Le Dilosquer, M., Singh, R. & Rycroft, M. J. Further considerations of engine emissions from subsonic aircraft at cruise altitude. Atmos. Environ.30, 3689–3695. 10.1016/1352-2310(96)00113-6 (1996). [Google Scholar]
  • 281.Neydorf, R., Krukhmalev, V., Kudinov, N. & Pshikhopov, V. Methods of Statistical Processing of Meteorological Data for the Tasks of Trajectory Planning of MAAT Feeders, 2013-01–2266 10.4271/2013-01-2266 (2013).
  • 282.Yonemura, S. et al. Temporal variations in ozone concentrations derived from Principal Component Analysis. Theor. Appl. Climatol.92, 47–58. 10.1007/s00704-007-0314-4 (2008). [Google Scholar]
  • 283.Gettelman, A. et al. The tropical tropopause layer 1960–2100. Atmos. Chem. Phys.9, 1621–1637. 10.5194/acp-9-1621-2009 (2009). [Google Scholar]
  • 284.Randel, W. J. & Jensen, E. J. Physical processes in the tropical tropopause layer and their roles in a changing climate. Nat. Geosci.6, 169–176. 10.1038/ngeo1733 (2013). [Google Scholar]
  • 285.Gettelman, A. & Forster, P. M. D. F. A climatology of the tropical tropopause layer. J. Meteorol. Soc. Jpn.80, 911–924. 10.2151/jmsj.80.911 (2002). [Google Scholar]
  • 286.Fueglistaler, S. et al. Tropical tropopause layer. Rev. Geophys.10.1029/2008RG000267 (2009). [Google Scholar]
  • 287.Randall, C. M. & Greenberg, L. T. Radiative transfer and 4.3 µm atmospheric clutter observations. In Modern Utilization of Infrared Technology VII, 126–133. 10.1117/12.932676 (SPIE, 1982).
  • 288.Mendoza, V. M., Villanueva, E. E., Garduño, R. & Sánchez-Meneses, O. Atmospheric emissivity with clear sky computed by E-Trans/HITRAN. Atmos. Environ.155, 174–188. 10.1016/j.atmosenv.2017.01.048 (2017). [Google Scholar]
  • 289.Marzouk, O. A. Subcritical and supercritical Rankine steam cycles, under elevated temperatures up to 900°C and absolute pressures up to 400 bara. Adv. Mech. Eng.16, 16878132231221064. 10.1177/16878132231221065 (2024). [Google Scholar]
  • 290.Kashkin, V. B., Vladimirov, V. M. & Klykov, A. O. Zenith tropospheric delay of GLONASS/GPS signals on the basis of ATOVS satellite data. Atmos. Ocean Opt.28, 68–73. 10.1134/S1024856015010066 (2015). [Google Scholar]
  • 291.Mohanakumar, K. Stratosphere Troposphere Interactions: An Introduction (Springer Science & Business Media, 2008). [Google Scholar]
  • 292.Essa, K. S. M., Etman, S. M. & Embaby, M. The effect of air density variation on wind power flux. Wind Eng.28, 305–310. 10.1260/0309524041590107 (2004). [Google Scholar]
  • 293.Marzouk, O. A. Benchmarks for the Omani higher education students-faculty ratio (SFR) based on World Bank data, QS rankings, and THE rankings. Cogent Educ.11, 2317117. 10.1080/2331186X.2024.2317117 (2024). [Google Scholar]
  • 294.Colás Irazusta, D. Study of numerical optimization of a flight trajectory, Bachelor thesis, http://hdl.handle.net/2117/414738 (Accessed 16 Dec 2024) (Universitat Politècnica de Catalunya, 2024).
  • 295.Marzouk, O. A. Changes in fluctuation waves in coherent airflow structures with input perturbation. WSEAS Trans. Signal Process.4, 604–614. 10.48550/arXiv.2410.08542 (2008). [Google Scholar]
  • 296.Marzouk, O. A. Investigation of Strouhal number effect on acoustic fields. In 22nd National Conference on Noise Control Engineering (NOISE-CON 2007),1056–1067. 10.31219/osf.io/38ehb (INCE [Institute of Noise Control Engineering], 2007).
  • 297.Wang, S., Xiao, S. & Tian, L. Penetration of oblique shock wave through different thickness boundary layers. Aerosp. Sci. Technol.157, 109863. 10.1016/j.ast.2024.109863 (2025). [Google Scholar]
  • 298.Yao, B. et al. Shock waves characteristics and losses estimation of non-equilibrium condensation flow in nozzle and steam turbine cascade. Appl. Therm. Eng.258, 124579. 10.1016/j.applthermaleng.2024.124579 (2025). [Google Scholar]
  • 299.Marzouk, O. A. Noise emissions from excited jets. In 22nd National Conference on Noise Control Engineering (NOISE-CON 2007), 1034–1045. 10.31219/osf.io/hpved (INCE [Institute of Noise Control Engineering], 2007).
  • 300.Nakatani, H., Taguchi, H., Fujita, K., Shindo, S. & Honami, S. Wind tunnel test of mach 5 class hypersonic airplane. Trans JSASS Space Technol. Jpn.7, Pe_71-Pe_76. 10.2322/tstj.7.Pe_71 (2009). [Google Scholar]
  • 301.Marzouk, O. A. A two-step computational aeroacoustics method applied to high-speed flows. Noise Control Eng. J.56, 396. 10.3397/1.2978229 (2008). [Google Scholar]
  • 302.Aspray, C. J. Design and analysis of a Mach5+ hypersonic wind tunnel, Master’s Thesis, United States Naval Postgraduate School. https://apps.dtic.mil/sti/pdfs/AD1150811.pdf (Accessed 17 Dec 2024) (2021).
  • 303.Marzouk, O. A. Detailed and simplified plasma models in combined-cycle magnetohydrodynamic power systems. Int. J. Adv. Appl. Sci.10, 96–108. 10.21833/ijaas.2023.11.013 (2023). [Google Scholar]
  • 304.Wie, D. M. V. Hypersonics: Past, Present, and Potential Future, Johns Hopkins APL Technical Digest 35 https://secwww.jhuapl.edu/techdigest/content/techdigest/pdf/v35-n04/35-04-van%20wie.pdf (Accessed 17 Dec 2024) (2021).
  • 305.Laney, C. B. Computational Gasdynamics 1st edn. (Cambridge University Press, 1998). 10.1017/CBO9780511605604. [Google Scholar]
  • 306.Marzouk, O. A. Hydrogen utilization as a plasma source for magnetohydrodynamic direct power extraction (MHD-DPE). IEEE Access12, 167088–167107. 10.1109/ACCESS.2024.3496796 (2024). [Google Scholar]
  • 307.Cebeci, T. ed. Computational Fluid Dynamics for Engineers: From Panel to Navier-Stokes Methods with Computer Programs (Horizons Pub. Inc. Springer, 2005).
  • 308.von Backstrom, T. W. The effect of specific heat ratio on the performance of compressible flow turbo-machines. In American Society of Mechanical Engineers Digital Collection, 2111–2117 10.1115/GT2008-50183 (2009).
  • 309.Marzouk, O. A. The Sod gasdynamics problem as a tool for benchmarking face flux construction in the finite volume method. Sci. Afr.10, e00573. 10.1016/j.sciaf.2020.e00573 (2020). [Google Scholar]
  • 310.Roberts, S. K. & Sjolander, S. A. Effect of the specific heat ratio on the aerodynamic performance of turbomachinery. J. Eng. Gas Turbines Power127, 773–780. 10.1115/1.1995767 (2002). [Google Scholar]
  • 311.Pajayakrit, P. & Tongsawang, K. Introducing climb-cruise. NKRAFA J. Sci. Technol.20, 93–107 (2024). [Google Scholar]
  • 312.Nasir, R. E. M. et al. Flight performance of various blended wing-body small UAV designs. Jurnal Teknologi Sci. Eng.10.11113/jt.v75.5214 (2015). [Google Scholar]
  • 313.Antcliff, K. R. & Capristan, F. M. Conceptual design of the parallel electric-gas architecture with synergistic utilization scheme (PEGASUS) concept. In 18th AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference, AIAA 2017-4001 10.2514/6.2017-4001 (American Institute of Aeronautics and Astronautics, 2017).
  • 314.Mori, R. Fuel-saving climb procedure by reduced thrust near top of climb. J. Aircr.57, 800–806. 10.2514/1.C035200 (2020). [Google Scholar]
  • 315.Takahashi, T. T. Climb performance of very high T/W airframes. In AIAA Aviation Forum and Ascend 2024, AIAA 2024-3914 10.2514/6.2024-3914 (AIAA [American Institute of Aeronautics and Astronautics], 2024).
  • 316.Mori, R. Fuel Saving by Gradual Climb Procedure, in: Seventh SESAR Innovation Days (SIDs), 1–8. https://www.sesarju.eu/sites/default/files/documents/sid/2017/SIDs_2017_paper_7.pdf (Accessed 17 Dec 2024) (European Organization for the Safety of Air Navigation (EUROCONTROL), 2017).
  • 317.Mori, R. Simulator experiment and real world implementation of reduced climb thrust fuel saving procedure. In 2022 IEEE/AIAA 41st Digital Avionics Systems Conference (DASC), 1–8 10.1109/DASC55683.2022.9925747 (2022).
  • 318.Rice, C. C. Validation of approach and climb-out times-in-mode for aircraft emissions computation. Transp. Res. Rec.2003, 79–82. 10.3141/1850-09 (1850). [Google Scholar]
  • 319.Files, D. S., Webb, J. T. & Pilmanis, A. A. Depressurization in military aircraft: rates, rapidity, and health effects for 1055 incidents. Aviat. Space Environ. Med.76, 523–529 (2005). [PubMed] [Google Scholar]
  • 320.Tong, K.-O., Warren, A. & Brown, J. Continuous descent approach procedure development for noise abatement tests at Louisville International Airport, KY. In AIAA’s 3rd Annual Aviation Technology, Integration, and Operations (ATIO) Forum, AIAA 2003-6772 10.2514/6.2003-6772 (AIAA [American Institute of Aeronautics and Astronautics], 2003).
  • 321.Luers, J. & Hairies, P. Heavy rain influence on airplane accidents. J. Aircr.20, 187–191. 10.2514/3.44850 (1983). [Google Scholar]
  • 322.Chan, P. W. & Krus, H. Validation of a crosswind change criterion for building induced airflow disturbances using a flight simulator: case studies at the Hong Kong International Airport. Meteorol. Appl.23, 742–748. 10.1002/met.1598 (2016). [Google Scholar]
  • 323.Arunachalam, S. & Varadappan, A. M. S. Effect of supply air failure on cabin pressure control system of a fighter aircraft. Int. J. Aeronaut. Space Sci.24, 570–580. 10.1007/s42405-022-00526-8 (2023). [Google Scholar]
  • 324.West, J. B. A strategy for in-flight measurements of physiology of pilots of high-performance fighter aircraft. J. Appl. Physiol.115, 145–149. 10.1152/japplphysiol.00094.2013 (2013). [DOI] [PubMed] [Google Scholar]
  • 325.Sathiyaseelan, A. & Selvan, A. M. Modeling and Simulation of a Fighter Aircraft Cabin Pressure Control System Using AMESim (SAE International, 2022). 10.4271/2022-28-0078. [Google Scholar]
  • 326.Tada, K. A Transformation formula for determining the second density virial coefficient from the second acoustic virial coefficient. Int. J. Thermophys.43, 64. 10.1007/s10765-022-02992-4 (2022). [Google Scholar]
  • 327.McBride, B. J., Zehe, M. J. & Gordon, S. NASA Glenn Coefficients for Calculating Thermodynamic Properties of Individual Species, https://ntrs.nasa.gov/api/citations/20020085330/downloads/20020085330.pdf (Accessed 11 Oct 2011) (NASA [United States National Aeronautics and Space Administration], 2002).
  • 328.Bahadori, A. & Vuthaluru, H. B. Estimation of air specific heat ratio at elevated pressures using simple predictive tool. Energy Convers. Manag.52, 1526–1532. 10.1016/j.enconman.2010.10.022 (2011). [Google Scholar]
  • 329.Liu, K., Chen, D., Serbin, S. & Patlaichuk, V. The features of GTU thermal calculation. In Gas Turbines Structural Properties, Operation Principles and Design Features (eds Liu, K.) 99–108 10.1007/978-981-99-0977-3_8 (Springer Nature, Singapore, 2023).
  • 330.Nariyuki, Y. Thermodynamic modeling of the solar wind plasma in the presence of envelope-modulated low-frequency Alfvén waves. Phys. Plasmas10.1063/1.5024016 (2018). [Google Scholar]
  • 331.Madu, K. E. & Atah, C. M. Effects of variation of ambient temperature in energy interactions in a closed system. Int. J. Progress. Res. Eng. Manag. Sci.4, 393–401 (2024). [Google Scholar]
  • 332.Marzouk, O. A. & Huckaby, E. D. Assessment of syngas kinetic models for the prediction of a turbulent nonpremixed flame. In Fall Meeting of the Eastern States Section of the Combustion Institute 2009, College Park, Maryland, USA, 726–751. 10.14293/PR2199.001370.v1 (2009).
  • 333.[United States National Aeronautics and Space Administration] NASA, Specific Heats - Calorically Imperfect Gas, https://www.grc.nasa.gov/www/BGH/realspec.html (Accessed 17 Dec 2024) (2021).
  • 334.Ubi Tutors. Air Properties Calculator, https://ubitutors.com/thermal-properties-of-air-at-a-given-temperature/ (Accessed 17 Dec 2024) (Ubi Tutors, 2023).
  • 335.Iserles, A. ed. Runge–Kutta methods. In A First Course in the Numerical Analysis of Differential Equations, 2nd ed, 33–52. 10.1017/CBO9780511995569.006 (Cambridge University Press, 2008).
  • 336.Marzouk, O. A. & Huckaby, E. D. A comparative study of eight finite-rate chemistry kinetics for CO/H2 combustion. Eng. Appl. Comput. Fluid Mech.4, 331–356. 10.1080/19942060.2010.11015322 (2010). [Google Scholar]
  • 337.Butcher, J. C. Numerical Methods for Ordinary Differential Equations 3rd edn. (Wiley, 2016). [Google Scholar]
  • 338.Press, W. H., Teukolsky, S. A., Flannery, B. P. & Vetterling, W. T. Numerical Recipes in FORTRAN 77: Volume 1, Volume 1 of Fortran Numerical Recipes: The Art of Scientific Computing, https://books.google.com.om/books?id=7vuNLcQhg8UC (Accessed 20 Dec 2024) (Cambridge University Press, 1992).
  • 339.Marzouk, O. A. A nonlinear ODE system for the unsteady hydrodynamic force - a new approach. World Acad. Sci. Eng. Technol.39, 948–962. 10.48550/arXiv.2410.13892 (2009). [Google Scholar]
  • 340.Vetterling, W. T. & Press, W. H. Numerical Recipes: Example Book C, https://books.google.com.om/books?id=4t-sybVuoqoC (Accessed 20 Dec 2024) (Cambridge University Press, 1992).
  • 341.Brincklow, J. R., Montgomery, Z. S. & Hunsaker, D. F. Controlling roll-yaw coupling with aileron placement and wing twist. In AIAA Scitech 2021 Forum. 10.2514/6.2021-0327 (American Institute of Aeronautics and Astronautics, VIRTUAL EVENT, 2021).
  • 342.Brincklow, J. R., Montgomery, Z. S. & Hunsaker, D. F. Controlling roll-yaw coupling with aileron and twist design. Aeronaut. J.10.1017/aer.2024.21 (2024). [Google Scholar]
  • 343.Jahnke, C. C. On the roll–coupling instabilities of high–performance aircraft. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.356, 2223–2239. 10.1098/rsta.1998.0271 (1998). [Google Scholar]
  • 344.Pedreiro, N., Rock, S. M., Celik, Z. Z. & Roberts, L. Roll-yaw control at high angle of attack by forebody tangential blowing. J. Aircr.35, 69–77. 10.2514/2.2261 (1998). [Google Scholar]
  • 345.Stengel, R. F. Flight Dynamics 2nd edn. (Princeton University Press, 2022). [Google Scholar]
  • 346.Chapman, C. H., Robb, C. & Paul, R. C. Investigation of roll-yaw coupling for fixed-wing unmanned air vehicles. In AIAA SCITECH 2024 Forum, AIAA 2024-2648 10.2514/6.2024-2648 (American Institute of Aeronautics and Astronautics, 2024).
  • 347.Fan, Y., Lutze, F. H. & Cliff, E. M. Time-optimal lateral maneuvers of an aircraft. J. Guid. Control. Dyn.18, 1106–1112. 10.2514/3.21511 (1995). [Google Scholar]
  • 348.Marzouk, O. A. & Nayfeh, A. H. Detailed characteristics of the resonating and non-resonating flows past a moving cylinder. In 49th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, AIAA 2008-2311. 10.2514/6.2008-2311 (AIAA [American Institute of Aeronautics and Astronautics], 2008).
  • 349.Cho, H. & Han, C. Effect of sideslip angle on the aerodynamic characteristics of a following aircraft in close formation flight. J. Mech. Sci. Technol.29, 3691–3698. 10.1007/s12206-015-0814-6 (2015). [Google Scholar]
  • 350.Stone Jr, R. W. Estimation of the maximum angle of sideslip for determination of vertical-tail loads in rolling maneuvers. https://ntrs.nasa.gov/api/citations/19930092166/downloads/19930092166.pdf (Accessed 18 Dec 2024) (NACA [United States National Advisory Committee for Aeronautics], 1953).
  • 351.Marzouk, O. & Nayfeh, A. Physical interpretation of the nonlinear phenomena in excited wakes. In 46th AIAA Aerospace Sciences Meeting and Exhibit, AIAA 2008-1304. 10.2514/6.2008-1304 (AIAA [American Institute of Aeronautics and Astronautics], 2008).
  • 352.Lerro, A., Brandl, A. & Gili, P. Model-free scheme for angle-of-attack and angle-of-sideslip estimation. J. Guid. Control. Dyn.44, 595–600. 10.2514/1.G005591 (2021). [Google Scholar]
  • 353.Marzouk, O. A. & Nayfeh, A. H. A study of the forces on an oscillating cylinder. In ASME 2007 26th International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2007), 741–752. 10.1115/OMAE2007-29163 (ASME [American Society of Mechanical Engineers], 2009)
  • 354.Hoffmann, K. A. & Chiang, S. T. Computational Fluid Dynamics – Vol. 1, 4 ed, 2. print (Engineering Education System, 2004).
  • 355.Marzouk, O. A. Directivity and noise propagation for supersonic free jets. In 46th AIAA Aerospace Sciences Meeting and Exhibit, AIAA 2008-23. 10.2514/6.2008-23 (AIAA [American Institute of Aeronautics and Astronautics], 2008)
  • 356.Fornberg, B. Classroom note: calculation of weights in finite difference formulas. SIAM Rev.40, 685–691. 10.1137/S0036144596322507 (1998). [Google Scholar]
  • 357.Milne-Thomson, L. M. The Calculus of Finite Differences (American Mathematical Soc., 2000). [Google Scholar]
  • 358.McAndrew, I. R., Navarro, E. & Brig, G. O. Low speed and high altitude flight: Unmanned flight at low speed in altitudes above 50,000ft. In 2017 8th International Conference on Mechanical and Aerospace Engineering (ICMAE), 556–560 10.1109/ICMAE.2017.8038708 (2017).
  • 359.Xie, L., Wei, Z., Ding, D., Zhang, Z. & Tang, A. Long and short term maneuver trajectory prediction of UCAV based on deep learning. IEEE Access9, 32321–32340. 10.1109/ACCESS.2021.3060783 (2021). [Google Scholar]
  • 360.Wei, Z. et al. A tactical maneuver trajectory prediction method using gate recurrent unit based on triangle search optimization with AdaBoost. In 2021 33rd Chinese Control and Decision Conference (CCDC), 2325–2330 10.1109/CCDC52312.2021.9602340 (2021).
  • 361.Lai, T., Yi, T.-H., Li, H.-N. & Fu, X. An explicit fourth-order Runge-Kutta method for dynamic force identification. Int. J. Str. Stab. Dyn.17, 1750120. 10.1142/S0219455417501206 (2017). [Google Scholar]
  • 362.Hassan, H. N. & Saleh, H. K. Fourier spectral methods for solving some nonlinear partial differential equations. Int. J. Open Probl. Comput. Sci. Math.6, 144–179 (2013). [Google Scholar]
  • 363.Liu, J.-G., Wang, C. & Johnston, H. A fourth order scheme for incompressible Boussinesq equations. J. Sci. Comput.18, 253–285. 10.1023/A:1021168924020 (2003). [Google Scholar]
  • 364.Zingg, D. & Chisholm, T. Runge-Kutta methods for linear problems. In 12th Computational Fluid Dynamics Conference, AIAA-95-1756-CP. 10.2514/6.1995-1756 (AIAA [American Institute of Aeronautics and Astronautics], 1995).
  • 365.I.A. Nweke, On the accuracy of the backward difference formula of order four and runge-kutta of order four in solving a first order non-linear ordinary differential equation, http://hdl.handle.net/20.500.12306/6669 (Accessed 21 Dec 2024) (Kampala International University, 2016).
  • 366.Dassault Aviation, Mirage III: origins, characteristics and performance data, Dassault Aviation, a Major Player to Aeronautics https://www.dassault-aviation.com/en/passion/aircraft/military-dassault-aircraft/mirage-iii (Accessed 21 Dec 2024) (2024).
  • 367.Dassault Aviation, Mirage III F2 F3: origins, characteristics and performance data, Dassault Aviation, a Major Player to Aeronautics. https://www.dassault-aviation.com/en/passion/aircraft/military-dassault-aircraft/mirage-iii-f2-f3/ (Accessed 21 Dec 2024) (2024).
  • 368.Aloni, S. Israeli Mirage III and Nesher Aces. https://books.google.com.om/books?id=T5weDAAAQBAJ (Bloomsbury Publishing, 2012).
  • 369.Aloni, S. Mirage III vs MiG-21: Six Day War 1967, https://books.google.com.om/books?id=LjS1CwAAQBAJ (Accessed 21 Dec 2024) (Bloomsbury Publishing, 2012).
  • 370.Deplante, H. The Mirage III–V V/STOL fighter, In: SAE World Congress & Exhibition, Detroit, Michigan, USA, 640217. 10.4271/640217 (1964).
  • 371.Dassault Aviation, Photo of Mirage III V 01, in flight, https://images.dassault-aviation.com/f_auto,q_auto,g_center,dpr_auto/wp-auto-upload/2/files/2018/06/DA00020244_Si.jpg (Accessed 18 Feb 2025) (2025).
  • 372.Eaton, J. W., Bateman, D., Hauberg, S. & Wehbring, R. GNU Octave version 6.1.0 manual: a high-level interactive language for numerical computations, https://www.gnu.org/software/octave/doc/v6.1.0 (Accessed 12 Feb 2025) (2020).
  • 373.Li, Y. & Huang, C. Configuration design of common aero vehicle based on genetic algorithm. In 2010 2nd International Conference on Information Engineering and Computer Science, 1–4 10.1109/ICIECS.2010.5678241 (2010).
  • 374.Huijuan, W. et al. A lightweight electromechanical actuation system with fast response. In 2016 IEEE International Conference on Aircraft Utility Systems (AUS), 596–599 10.1109/AUS.2016.7748121 (2016).
  • 375.Long, H., Wang, H., Zhang, H., Guo, X. & Wang, L. Research on a lightweight integrated four-channel electromechanical actuation system. In CSAA/IET International Conference on Aircraft Utility Systems (AUS 2018), 1–5. 10.1049/cp.2018.0163 (2018).
  • 376.Clark, R. A., Tai, J. C. & Mavris, D. N. Development of a parametric variable cycle engine model using the multiple design point approach. In AIAA SCITECH 2023 Forum, AIAA 2023-1464. 10.2514/6.2023-1464 (AIAA [American Institute of Aeronautics and Astronautics], 2023).
  • 377.Marzouk, O. A. & Nayfeh, A. H. A parametric study and optimization of ship-stabilization systems. In 1st WSEAS International Conference on Maritime and Naval Science and Engineering (MN’08), 169–174. https://doi.org/10.31219/osf.io/9w3gc_v1 (WSEAS [World Scientific and Engineering Academy and Society], 2008).
  • 378.White, M. & Padfield, G. The use of flight simulation for research and teaching in academia. In AIAA Atmospheric Flight Mechanics Conference and Exhibit, AIAA 2006-6493. 10.2514/6.2006-6493 (AIAA [American Institute of Aeronautics and Astronautics], 2006).
  • 379.Marzouk, O. A. Benchmarking retention, progression, and graduation rates in undergraduate higher education across different time windows. Cogent Educ.12, 2498170. 10.1080/2331186X.2025.2498170 (2025). [Google Scholar]
  • 380.Gohardani, O., Gohardani, A. S., Dokter, E. & Macario, K. Aeronautical engineering and aerospace engineering: a learner-centered teaching perspective in higher education. J. Coll. Sci. Teach.44, 64–71. 10.2505/4/jcst14_044_01_64 (2014). [Google Scholar]
  • 381.Marzouk, O. A. Accrediting artificial intelligence programs from the omani and the international ABET perspectives. In Intelligent Computing (ed Arai, K.), 462–474. 10.1007/978-3-030-80129-8_33 (Springer International Publishing, 2021).
  • 382.Celedón-Pattichis, S. et al. Asset-Based Approaches to Equitable Mathematics Education Research and Practice10.5951/jresematheduc.49.4.0373 (2018).
  • 383.Stevens, B. L., Lewis, F. L. & Johnson, E. N. Aircraft Control and Simulation: Dynamics, Controls Design, and Autonomous Systems: Dynamics, Controls Design, and Autonomous Systems, 1st ed. 10.1002/9781119174882 (Wiley, 2015).
  • 384.Steger, C. R., Steger, B. & Schär, C. HORAYZON v12: An efficient and flexible ray-tracing algorithm to compute horizon and sky view factor. Geosci. Model Dev.15, 6817–6840. 10.5194/gmd-15-6817-2022 (2022). [Google Scholar]
  • 385.Krause, L. H. et al. Science conops for application of sport mission data to study large (~1000KN) ionospheric plasma depletions. In 36th Annual Small Satellite Conference, 1–14. https://ntrs.nasa.gov/citations/20220012007 (Accessed 17 May 2025) (USU [Utah State University], 2022).
  • 386.MathWorks, MATLAB Simulink (Aerospace Blockset) │ 6DOF ECEF (Quaternion) - Implement quaternion representation of six-degrees-of-freedom equations of motion in Earth-centered Earth-fixed (ECEF) coordinates. https://www.mathworks.com/help/aeroblks/6dofecefquaternion.html (Accessed 17 May 2025) (2025).
  • 387.Kisabo, A. B., Adebimpe, A. F., Kisabo, A. B. & Adebimpe, A. F. State-space modeling of a rocket for optimal control system design. In Ballistics, 10.5772/intechopen.82292 (IntechOpen, 2019).
  • 388.Bayiz, Y. E. & Cheng, B. State-space aerodynamic model reveals high force control authority and predictability in flapping flight. J. R. Soc. Interface18, 20210222. 10.1098/rsif.2021.0222 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.MathWorks, MATLAB │ Analyze State-Space Model for Linear Control and Static Stability Analysis, https://ww2.mathworks.cn/help/aerotbx/ug/perform-controls-and-static-stability-analysis-linearized-fixed-wing-aircraft.html (Accessed 16 May 2025) (2025).
  • 390.Hemingway, E. G. & O’Reilly, O. M. Perspectives on Euler angle singularities, gimbal lock, and the orthogonality of applied forces and applied moments. Multibody Syst. Dyn.44, 31–56. 10.1007/s11044-018-9620-0 (2018). [Google Scholar]
  • 391.Brezov, D. S., Mladenova, C. D. & Mladenov, I. M. New perspective on the gimbal lock problem. AIP Conf. Proc.1570, 367–374. 10.1063/1.4854778 (2013). [Google Scholar]
  • 392.Maine, R. E. & Iliff, K. W. Application of parameter estimation to aircraft stability and control: the output-error approach. https://ntrs.nasa.gov/api/citations/19870020066/downloads/19870020066.pdf (Accessed 3 Apr 2024) (NASA [United States National Aeronautics and Space Administration], 1986).
  • 393.Mansur, V., Reddy, S. & Sujatha, R. Deploying complementary filter to avert gimbal lock in drones using quaternion angles. In 2020 IEEE International Conference on Computing, Power and Communication Technologies (GUCON), 751–756 10.1109/GUCON48875.2020.9231126 (2020).
  • 394.Kang, C. W. & Park, C. G. Euler angle based attitude estimation avoiding the singularity problem. IFAC Proc. Vol.44, 2096–2102. 10.3182/20110828-6-IT-1002.01993 (2011). [Google Scholar]
  • 395.Bernardes, E. & Viollet, S. Quaternion to Euler angles conversion: A direct, general and computationally efficient method. PLoS ONE17, e0276302. 10.1371/journal.pone.0276302 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

Data generated or analyzed during this study are included in this published article. A supplementary material containing expressions for second time derivatives that were not presented in the main body of the manuscript is provided as a document titled (suppl.pdf).


Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES