Fact-checked by Grok 2 weeks ago

Euler angles

Euler angles are three angular parameters, typically denoted as φ, θ, and ψ, that specify the orientation of a in relative to a fixed through a sequence of three successive rotations about designated axes. Introduced by the Leonhard Euler in his 1776 paper Formulae generales pro translatione quacunque corporum rigidorum, these angles provide a parameterization of rotations based on , which states that any displacement of a can be described by a single rotation about an axis passing through the body's fixed point. The standard formulation involves composing rotation matrices, such as R = R_z(ψ) R_y(θ) R_x(φ) for the common z-y-x convention, where each matrix represents an elementary rotation about the respective axis. Various conventions exist for the order of rotations, including the Euler convention (e.g., z-x-z sequence) and Tait-Bryan angles (e.g., z-y-x, often called yaw-pitch-roll in applications), each suited to specific fields like , , and . These angles are particularly useful for expressing the of , aircraft, and gyroscopes, where φ might represent , θ , and ψ spin. The and of a rotating can be formulated in terms of Euler angles and their time derivatives, facilitating the derivation of Euler's for torque-free rotation. Despite their intuitiveness, Euler angles exhibit mathematical singularities known as , occurring when one of the intermediate angles reaches 0 or π, causing two rotation axes to align and reducing the three degrees of freedom to two. For instance, in the z-y-x sequence, when θ = π/2, the effective rotation about the y-axis aligns the z and x axes, making independent adjustments of φ and ψ indistinguishable in certain directions. This limitation has prompted alternatives like quaternions for in applications such as flight simulation and 3D animation.

Fundamentals

Definition and historical context

Euler angles constitute a three-parameter representation of the orientation of in three-dimensional relative to a fixed . These , conventionally denoted as \alpha, \beta, and \gamma, describe the body's through a composition of three successive rotations about specified coordinate axes, enabling the transformation from the to the body frame. This parameterization draws on the geometric insight that any orientation in the special orthogonal group SO(3)—the Lie group of all proper rotations in 3D space—can be achieved via such sequential rotations around orthogonal axes, as established by Euler's foundational work on displacements. A example of this approach in proper Euler angles employs the z-x-z sequence: the first rotation by \alpha about the initial z-axis, followed by a by \beta about the intermediate x-axis, and concluding with a by \gamma about the final z-axis. This method provides an intuitive, human-interpretable way to quantify and manipulate orientations, particularly in fields requiring precise control, though it is distinct from Tait-Bryan angles, which instead sequence rotations about three mutually perpendicular axes. Leonhard Euler first introduced the concept of these three angles in his investigations into rigid body motion during the 1770s, culminating in the 1776 publication "Formulae generales pro translatione quacunque corporum rigidorum" in the Novi Commentarii Academiae Scientiarum Petropolitanae. In this seminal paper, Euler demonstrated that arbitrary finite displacements of a , preserving a fixed point, equate to a single about an axis through that point, and he parameterized the general using direction cosines and angles, laying the groundwork for the successive framework. Earlier related ideas appeared in his paper "De motu corporum circa punctum fixum mobilium" (written after 1751, published posthumously in 1862) and the 1760 paper "Du mouvement d’un corps solide quelconque lorsqu’il tourne autour d’un axe mobile," where he explored components via three angles for symmetric bodies. The development of Euler angles continued to evolve in the , with significant advancements by French Benjamin , who in 1840 published explicit formulas for composing successive finite rotations and introduced parameters (now known as Euler-Rodrigues parameters) that complemented the angular description. Concurrently, Irish William advanced rotation theory through his 1843-1844 invention of quaternions, a four-dimensional algebra that offered an alternative, non-singular parameterization of SO(3) and influenced subsequent refinements to Euler's angular methods. These contributions solidified Euler angles as a cornerstone of 3D rotation kinematics, bridging with modern group-theoretic interpretations.

Relation to rotations in 3D space

Euler's rotation theorem asserts that any orientation of a in three-dimensional can be achieved by a single about some fixed axis passing through a point, such as the origin. This theorem underpins the use of Euler angles, which provide an alternative representation by decomposing the same overall into a sequence of three successive rotations about specific axes, either fixed in space or attached to the body. Such decompositions are possible because the composition of three rotations suffices to span the full space of possible orientations, as rotations form a three-dimensional configuration space. The equivalence between intrinsic and extrinsic rotation sequences arises from the properties of rotation matrix multiplication. In the extrinsic case, rotations are applied successively about fixed axes in the reference frame, resulting in a matrix product R = R_3 R_2 R_1, where each R_i is a about a fixed . For intrinsic rotations, the axes rotate with the , leading to R = R_1 R_2 R_3, but with the axes updated after each step; this is mathematically equivalent to the extrinsic product with the order of rotations reversed, i.e., R_1 R_2 R_3 = R_3' R_2' R_1', where the primed matrices correspond to the reversed sequence. This reversal ensures that both approaches yield the identical net orientation, as the non-commutativity of rotations is preserved through the adjusted order. In the context of , Euler angles serve as local coordinates on the special SO(3), which parameterizes all proper rotations in three dimensions as 3×3 orthogonal matrices with determinant 1. Unlike a , SO(3) is a non-commutative , meaning that the order of composing two rotations affects the result—rotating first around one and then another does not commute with the reverse order—necessitating careful sequencing in Euler angle representations. To illustrate, consider orienting an object like a rigid : the first might tilt it away from its initial alignment, the second adjusts the tilt direction, and the third spins it around the final axis, with the cumulative effect matching a direct axis-angle but distributed across the three steps for intuitive decomposition.

Proper Euler Angles

Intrinsic

Proper Euler angles, also known as classic Euler angles, represent an orientation in through a sequence of three successive about axes that are fixed relative to the rotating body itself. This intrinsic approach contrasts with extrinsic rotations by updating the rotation axes after each step, ensuring that subsequent rotations occur in the body's local coordinate frame. Common sequences for proper Euler angles include z-x-z, z-y-z, and x-y-z, where the first and third axes coincide, allowing the parameterization to cover the full special orthogonal group SO(3) except for certain singular configurations. Consider the z-x-z sequence as a representative example. The composition begins with an initial by \alpha () around the body's z-, denoted as R_z(\alpha). This is followed by a second by \beta () around the newly aligned x'-, represented by R_{x'}(\beta). Finally, a third by \gamma (intrinsic or ) occurs around the updated z''-, given by R_{z''}(\gamma). The overall R is the product of these individual matrices in the order of application: R = R_z(\gamma) R_x(\beta) R_z(\alpha), where the matrices are defined in the body-fixed frame and the reflects the intrinsic nature of the sequence. The explicit form of the rotation matrices for the elemental rotations are standard in three dimensions: R_z(\alpha) = \begin{pmatrix} \cos\alpha & -\sin\alpha & 0 \\ \sin\alpha & \cos\alpha & 0 \\ 0 & 0 & 1 \end{pmatrix}, \quad R_x(\beta) = \begin{pmatrix} 1 & 0 & 0 \\ 0 & \cos\beta & -\sin\beta \\ 0 & \sin\beta & \cos\beta \end{pmatrix}. Multiplying these yields the composite z-x-z matrix: R = \begin{pmatrix} \cos\alpha \cos\beta \cos\gamma - \sin\alpha \sin\gamma & -\cos\alpha \cos\beta \sin\gamma - \sin\alpha \cos\gamma & \cos\alpha \sin\beta \\ \sin\alpha \cos\beta \cos\gamma + \cos\alpha \sin\gamma & -\sin\alpha \cos\beta \sin\gamma + \cos\alpha \cos\gamma & \sin\alpha \sin\beta \\ -\sin\beta \cos\gamma & \sin\beta \sin\gamma & \cos\beta \end{pmatrix}. This formulation assumes right-handed rotations with positive angles following the . Geometrically, the intrinsic sequence implies that the coordinate axes co-rotate with the body, transforming the rotation into a path on the unit sphere where each step adjusts the local . This body-fixed perspective leads to non-commutativity of the rotations, meaning the order of application affects the final , as is not commutative; for instance, swapping the order of \alpha and \gamma in the z-x-z sequence produces a different result. Such sequences are particularly useful in applications like where tracking local orientations is essential.

Extrinsic rotation sequence

In the extrinsic formulation of proper Euler angles, the orientation of a rigid body is described by a sequence of three successive rotations performed around fixed axes in the reference (space-fixed) coordinate frame, rather than axes attached to the body itself. This approach composes the rotations in a straightforward manner using the laboratory or world coordinate system, making it particularly suitable for scenarios where the reference frame remains stationary, such as in certain astronomical or simulation contexts. Unlike body-fixed rotations, each extrinsic rotation operates on the current orientation relative to the unchanging external axes, ensuring that the sequence directly accumulates transformations in the global frame. A canonical example is the z-x-z extrinsic sequence, where the first rotation is by angle γ around the fixed z-axis, followed by a rotation by β around the fixed x-axis, and finally a by α around the fixed z-axis. The resulting total matrix R is obtained by multiplying the individual elementary matrices in the order of application, from right to left: \mathbf{R} = \mathbf{R}_z(\alpha) \mathbf{R}_x(\beta) \mathbf{R}_z(\gamma), where \mathbf{R}_z(\gamma) = \begin{pmatrix} \cos\gamma & -\sin\gamma & 0 \\ \sin\gamma & \cos\gamma & 0 \\ 0 & 0 & 1 \end{pmatrix}, \quad \mathbf{R}_x(\beta) = \begin{pmatrix} 1 & 0 & 0 \\ 0 & \cos\beta & -\sin\beta \\ 0 & \sin\beta & \cos\beta \end{pmatrix}, \quad \mathbf{R}_z(\alpha) = \begin{pmatrix} \cos\alpha & -\sin\alpha & 0 \\ \sin\alpha & \cos\alpha & 0 \\ 0 & 0 & 1 \end{pmatrix}. This composition yields a 3×3 that maps vectors from the body frame to the reference frame, preserving the right-handed convention. The extrinsic z-x-z is mathematically equivalent to the corresponding intrinsic (body-fixed) but with the first and third interchanged (α ↔ γ) and the overall of application reversed, due to the property of matrices where the of a is its . This duality arises because composing around fixed axes in one mirrors the effect of body-fixed in the reverse , allowing the same to be represented interchangeably under the two paradigms. One advantage of the extrinsic formulation lies in its computational simplicity for fixed-frame analyses, such as in or , where transformations can be applied directly without updating intermediate axis orientations after each step. This avoids the need to track rotating frames, reducing complexity in implementations that prioritize global coordinate consistency over body-centric interpretations.

Conventions, signs, and ranges

In proper Euler angles, the sign convention for rotations adheres to the , where a positive rotation about an axis is defined such that the thumb of the right hand points along the positive direction of the and the fingers curl in the direction of the . This rule ensures consistency in defining the sense of rotation, with positive angles corresponding to counterclockwise motion when looking along the from the positive end, equivalent to a right-hand advancing into the positive direction. Interpretations can be active, where rotations transform coordinates from a fixed to a body (intrinsic sequence), or passive, where they describe reorientations without moving the body (extrinsic sequence), though the resulting orientation matrix remains the same for equivalent sequences. For the common z-x-z (or 3-1-3) sequence in proper Euler angles, typical ranges are chosen to cover all orientations uniquely while minimizing redundancy: the first angle α (or φ) spans 0 to 2π, the second angle β (or θ) spans 0 to π, and the third angle γ (or ψ) spans 0 to 2π. These limits prevent overlap in most cases, as β exceeding π would duplicate orientations via , and the 2π periodicity of α and γ accounts for full . In practice, software and fields like cryo-electron microscopy enforce these ranges to standardize computations. Notations for proper Euler angles vary between historical and modern usages. Leonhard Euler originally employed Greek letters α, β, γ for the angles in his 1776 work on motion, with α as the initial rotation about the z-axis, β about the line of nodes (x'-axis in z-x-z), and γ about the final z''-axis. Contemporary physics and often use φ for (first angle), θ for (second), and ψ for (third), particularly in the z-x-z convention prevalent in textbooks. In applications, the 3-1-3 sequence (z-x-z) is occasionally used with φ, θ, ψ notation, though the more asymmetric (z-y-x) Tait-Bryan sequence dominates for vehicle attitudes. A key challenge with proper Euler angles is their non-uniqueness: the same can correspond to multiple angle triples due to the periodicity and symmetries of rotations. For instance, in the z-x-z convention, (α, β, γ) represents the same as (α + 2π, β, γ) or (α, 2π - β, -γ), and at β = 0 or β = π ( points), infinitely many combinations arise since the intermediate axis aligns with the others. Additionally, (α + π, π - β, γ + π) yields an equivalent rotation, reflecting the SO(3) group's properties. These ambiguities are resolved through the standard range restrictions and field-specific conventions, ensuring a set for each .

Precession, nutation, and spin

In the z-x-z convention of proper Euler angles, the angles carry specific physical interpretations related to the components of a rigid body's . The first angle, α, represents the , which is the of the body's symmetry around the fixed space z-. The second angle, β, denotes the , corresponding to the tilt or nodding motion of the symmetry away from the fixed z-. Finally, the third angle, γ, describes the intrinsic , which is the of the body about its own symmetry . These interpretations are particularly apt for symmetric rigid bodies, such as or , where the z- aligns with the principal of inertia. The instantaneous angular velocity vector ω of the body can be decomposed as the vector sum of contributions from each angle's time derivative, expressed in terms of the evolving reference frames: \boldsymbol{\omega} = \dot{\alpha} \mathbf{k} + \dot{\beta} \mathbf{i}' + \dot{\gamma} \mathbf{k}'' Here, \mathbf{k} is the unit vector along the fixed space z-axis, \mathbf{i}' is the unit vector along the intermediate x-axis after the rotation, and \mathbf{k}'' is the unit vector along the body's z-axis after all . This decomposition highlights how the total arises from the superposition of precessional, nutational, and motions, with each term aligned to its respective at the instant considered. In , this decomposition facilitates the application of Euler's , which describe the evolution of ω under torques. For a torque-free symmetric , the equations simplify to show that the spin component \dot{\gamma} remains constant, while and couple to produce polhode motion on the 's energy . With external torques, such as on an , the equations reveal steady solutions where \dot{\alpha} balances the torque-induced wobble, linking the Euler angles directly to analyses in systems like gyroscopes. A prominent example is the rotation of Earth, modeled using Euler angles to capture its complex orientation. The daily spin corresponds to rapid changes in γ, with a period of approximately 24 hours. Precession manifests as the slow westward drift of the equinoxes around the ecliptic pole, driven by gravitational torques from the Sun and Moon on Earth's equatorial bulge, completing a cycle every 25,800 years. Nutation superimposes small oscillatory tilts in β, primarily due to the Moon's orbital inclination and nodal precession, with principal amplitudes of about 9.2 arcseconds in latitude and 17.2 arcseconds in longitude over an 18.6-year period. This framework, rooted in Euler's original analysis, underpins modern celestial mechanics for predicting Earth's orientation parameters.

Tait-Bryan Angles

Definitions and sequences

Tait-Bryan angles represent the orientation of a in through a sequence of three successive rotations about distinct axes, typically chosen from the orthogonal triad {x, y, z}. These angles, also referred to as Cardan angles, form an asymmetric set in contrast to the symmetric proper Euler angles, which repeat the first and third rotation axes. There are six possible Tait-Bryan sequences, corresponding to the six possible permutations of the three axes: 1-2-3 (XYZ), 1-3-2 (XZY), 2-1-3 (YXZ), 2-3-1 (YZX), 3-1-2 (ZXY), and 3-2-1 (ZYX). A widely used example is the ZYX sequence, which applies a yaw rotation about the z-axis, followed by a pitch rotation about the intermediate y-axis, and a roll rotation about the final x-axis. In the intrinsic formulation of the ZYX sequence, where each rotation is performed about the body-fixed axes updated after the previous rotation, the composite is composed as \mathbf{R} = \mathbf{R}_x(\gamma) \mathbf{R}_y(\beta) \mathbf{R}_z(\alpha), where \alpha denotes the yaw angle, \beta the pitch angle, \gamma the roll angle, and \mathbf{R}_x, \mathbf{R}_y, \mathbf{R}_z are the standard rotation matrices about the x-, y-, and z-axes, respectively. These distinct-axis sequences enable a full parameterization of SO(3), the special orthogonal group of 3D s, but they exhibit geometric asymmetries, such as restricted ranges for the intermediate angle to ensure unique representations and avoid singularities—for instance, in ZYX, is confined to (-\pi/2, \pi/2) while yaw and span [0, 2\pi).

Common conventions

In and , the most common Tait-Bryan convention employs the ZYX intrinsic sequence, where the angles are yaw (ψ) about the z-, (θ) about the y-, and roll (φ) about the x-, applied sequentially in the body-fixed frame. This sequence begins with a yaw about the initial vertical , followed by a about the intermediate lateral , and concludes with a roll about the final longitudinal . The corresponding transforming from the inertial frame to the body frame is given by the product \mathbf{H}_I^B = \mathbf{H}_x(\phi) \mathbf{H}_y(\theta) \mathbf{H}_z(\psi), where each \mathbf{H} denotes a basic rotation matrix around the respective axis. In nautical applications, the same ZYX intrinsic sequence is widely adopted to describe vessel orientation, with the yaw angle ψ often termed "heading" to denote the direction relative to north, while pitch and roll retain their standard meanings for vertical and transverse motions. This convention facilitates the representation of a ship's attitude in terms of its course (heading), trim (pitch), and list (roll). To mitigate singularities such as , typical ranges for these angles are restricted to ψ ∈ [0, 2π), θ ∈ [-π/2, π/2], and φ ∈ [0, 2π), ensuring the pitch angle avoids alignment that couples yaw and roll. Alternative formulations distinguish between intrinsic (body-fixed axes) and extrinsic (space-fixed axes) rotations. For the extrinsic ZYX equivalent, the sequence applies rotations in the reverse order about fixed axes, yielding the matrix product \mathbf{H}_I^B = \mathbf{H}_z(\psi) \mathbf{H}_y(\theta) \mathbf{H}_x(\phi), which achieves the same overall but differs in intermediate frames. This duality allows flexibility in computational implementations while preserving the final attitude description in both fields.

Alternative names and equivalences

Tait-Bryan angles are historically known as Cardan angles, named after the polymath (1501–1576), who described the use of gimbals to maintain orientation in his 1550 work De subtilitate rerum. They are also referred to as Bryan angles, honoring the contributions of British mathematician George Hartley Bryan (1864–1928) to the of flight and rotation theory in the late . In modern and , these angles are commonly called roll-pitch-yaw (RPY) angles, where roll denotes rotation about the forward axis, pitch about the lateral axis, and yaw about the vertical axis, providing an intuitive parameterization for asymmetric body orientations. Tait-Bryan angles form a subclass of the broader family of Euler angle parameterizations, distinguished by their use of three distinct axes (e.g., z-y-x) rather than the repeated axes characteristic of proper Euler angles (e.g., z-x-z). This asymmetry makes Tait-Bryan angles equivalent to proper Euler angles in the limit where the intermediate angle approaches specific values that align the effective axes, though both share the same underlying SO(3) manifold structure. Unlike proper Euler angles, which preserve around a principal , Tait-Bryan sequences avoid inherent axis repetition, facilitating distinct mathematical treatments in non-symmetric contexts. Tait-Bryan angles are favored for applications requiring intuitive descriptions of or attitudes, such as in where yaw, , and roll directly map to heading, , and banking maneuvers. In contrast, proper Euler angles are typically employed in problems involving symmetric rigid bodies, like the torque-free motion of a spinning , where the repeated aligns with the body's to simplify formulations. The following table summarizes common Tait-Bryan sequences, their conventional names, and typical domains of use:
Sequence (Intrinsic)AnglesCommon NameTypical Use Case
z-y'-x''Yaw (ψ), Pitch (θ), Roll (φ)Yaw-Pitch-Roll, mobile robotics
x-y'-z''Roll (φ), Pitch (θ), Yaw (ψ)Roll-Pitch-Yaw, some manipulators
z-y'-x''Yaw (ψ), Pitch (θ), Roll (φ)Yaw-Pitch-RollNautical

Extracting Angles from Orientations

Proper Euler angles for a frame

Proper Euler angles describe the orientation of using a symmetric sequence of rotations, such as z-x-z, where the initial and final rotations are about the same axis type (z-axis) and the intermediate rotation is about a perpendicular axis (x-axis). To compute these angles from a given orientation, the frame's attitude is typically represented by a rotation matrix R, a 3×3 orthogonal matrix satisfying R^T R = I and \det(R) = 1, which ensures it corresponds to a proper rotation without reflections. The extraction process relies on equating elements of R to the trigonometric expressions derived from the composition of the individual rotation matrices in the sequence. Note that there may be two equivalent sets of angles representing the same rotation, differing by complementary intermediate angles and adjustments to the others. The algorithm for the z-x-z sequence begins by identifying the middle angle \beta, the rotation about the x-axis, using \beta = \arccos(R_{33}), where \beta \in [0, \pi]. This follows from the (3,3) element of the composite rotation matrix being \cos \beta. Next, the precession angle \alpha, the initial rotation about the z-axis, is found with \alpha = \atantwo(R_{31}, R_{32}), leveraging the two-argument arctangent to resolve quadrant ambiguities and yield \alpha \in (-\pi, \pi]; here, the relevant elements satisfy R_{31} = \sin \beta \sin \alpha and R_{32} = \sin \beta \cos \alpha. Similarly, the spin angle \gamma, the final rotation about the new z-axis, is computed as \gamma = \atantwo(R_{13}, R_{23}), with \gamma \in (-\pi, \pi], based on elements where R_{13} = \sin \gamma \sin \beta and R_{23} = \cos \gamma \sin \beta in the standard convention. These steps assume \sin \beta \neq 0; the atan2 function inherently handles the signs and ranges for full coverage of the orientation space. As a numerical example, consider a sample with relevant elements R_{33} = 0.5, R_{31} \approx 0.612, R_{32} \approx 0.612, R_{13} \approx 0.612, and R_{23} \approx 0.612 (derived from \alpha = \pi/4, \beta = \pi/3, \gamma = \pi/4, with \sin(\pi/3) \approx 0.866, \cos(\pi/3) = 0.5, and \sin(\pi/4) = \cos(\pi/4) \approx 0.707). Then, \beta = \arccos(0.5) \approx 1.047 radians (60°). Next, \alpha = \atantwo(0.612, 0.612) \approx 0.785 radians (45°). Finally, \gamma = \atantwo(0.612, 0.612) \approx 0.785 radians (45°). This computation verifies the angles reconstruct the original orientation when composed into the .

Tait-Bryan angles for a frame

Tait-Bryan angles represent orientations using rotations about three distinct principal axes, commonly applied to describe the attitude of a frame such as an aircraft's body relative to a frame. In the ZYX sequence—prevalent in for its alignment with intuitive yaw (about Z), pitch (about Y'), and roll (about X'') motions—the angles are extracted from the direction cosine matrix () R that transforms vectors from the body frame to the reference frame. This sequence is particularly suited for frames where the pitch axis remains nearly aligned with the horizontal plane during typical maneuvers. Note that there may be two equivalent sets of angles representing the same rotation, differing by complementary intermediate angles and adjustments to the others. The extraction algorithm begins by computing the pitch angle \theta from the (3,1) element of R, which isolates the sine of the pitch due to the structure of the composite rotation matrix. Specifically, \theta = -\asin(R_{31}), where R_{31} is the element in the third row and first column. With \theta determined (assuming |\theta| < \pi/2 to avoid singularities), the yaw angle \psi is then found using the third row elements: \psi = \atan2(R_{32} / \cos \theta, R_{33} / \cos \theta). Finally, the roll angle \phi uses the first row elements: \phi = \atan2(R_{21} / \cos \theta, R_{11} / \cos \theta). These steps leverage the atan2 function to resolve the correct quadrant for each angle, ensuring the full range of [-\pi, \pi) for \psi and \phi, and [-\pi/2, \pi/2] for \theta. For an illustrative computation in an aircraft attitude matrix, consider the rotation matrix R = \begin{pmatrix} 0.5 & -0.1464 & 0.8536 \\ 0.5 & 0.8536 & -0.1464 \\ -0.7071 & 0.5 & 0.5 \end{pmatrix}, which might arise from a coordinated turn. Here, \theta = -\asin(-0.7071) = \pi/4 radians (approximately 45° pitch up). Then, \psi = \atan2(0.5 / \cos(\pi/4), 0.5 / \cos(\pi/4)) = \pi/4 radians (45° yaw), and \phi = \atan2(0.5 / \cos(\pi/4), 0.5 / \cos(\pi/4)) = \pi/4 radians (45° roll). This yields an orientation interpretable as a steady banked turn in aircraft dynamics. Unlike proper Euler angle extractions, which employ symmetric axis sequences (e.g., ZXZ) and are better suited for general symmetric orientations, Tait-Bryan methods like ZYX provide more straightforward computations for frames with near-vertical attitudes, as the distinct axes align directly with vehicle-specific motions such as heading, elevation, and bank.

Handling singularities

Singularities in the representation of orientations using Euler angles, commonly referred to as gimbal lock, arise when the intermediate rotation angle causes two of the rotation axes to align, resulting in a loss of one degree of freedom in the three-dimensional rotation space. For proper Euler angles, such as in the z-x-z sequence, this occurs when the intermediate angle β equals 0 or π radians, where the first and third rotation axes become collinear. In Tait-Bryan angles, like the common yaw-pitch-roll (z-y-x) convention, gimbal lock happens when the pitch angle θ reaches ±π/2 radians, aligning the yaw and roll axes. These conditions stem from the intrinsic topology of the special orthogonal group SO(3), which cannot be continuously parameterized by three angles without singularities. The primary effects of these singularities include ambiguity in the angle values and infinite possible solutions for the affected angles, leading to discontinuities or numerical instabilities in computations involving orientation extraction or interpolation. For instance, near the singular configuration, small changes in the rotation matrix can produce arbitrarily large jumps in the extracted angles, complicating algorithms for attitude determination in applications like aerospace. This loss of a degree of freedom means that rotations around the aligned axes cannot be uniquely decomposed, potentially causing errors in tracking or control systems. To handle these singularities during angle extraction from a rotation matrix, specialized algorithms detect the condition—typically by checking if the sine of the intermediate angle is zero or near-zero—and apply corrective measures. One common approach is to use complementary angles: for example, if θ ≈ π/2 in a , replace it with π - θ and adjust the adjacent angles by adding or subtracting π to maintain the equivalent rotation. Another method involves switching to an alternative Euler angle sequence that avoids the singularity at the current orientation, such as converting from a z-y-x to an x-y-z sequence via intermediate representations, ensuring continuity across the configuration space. In cases of exact singularity, the extraction formulas degenerate, and one angle is conventionally set to zero while solving for the combined effect in the others using arctangent functions on matrix elements. A practical example occurs in the z-y-x (yaw-pitch-roll) Tait-Bryan sequence, where gimbal lock at θ = ±π/2 renders yaw (ψ) and roll (φ) indeterminate from the standard extraction equations involving the rotation matrix elements r_{12} and r_{13}. Recovery involves setting ψ = 0 and computing φ as the arctangent of the ratio -r_{12}/r_{13} for θ = -π/2, or adjusting for the difference φ - ψ in the positive case, preserving the overall orientation. For broader avoidance, especially in real-time systems, quaternions offer a singularity-free alternative, though brief conversions back to angles may still require singularity-aware handling.

Mathematical Conversions

To and from rotation matrices

The conversion between Euler angles and rotation matrices is fundamental for representing three-dimensional orientations in a fixed coordinate system. For proper Euler angles, which involve rotations about the same axis type for the first and third angles, the rotation matrix is constructed as the product of three individual rotation matrices. A common convention is the z-x-z sequence, where the angles are denoted as α (first rotation about z), β (second about x), and γ (third about z). The overall rotation matrix R is given by R = R_z(\alpha) R_x(\beta) R_z(\gamma), where the individual matrices are: R_z(\alpha) = \begin{pmatrix} \cos\alpha & \sin\alpha & 0 \\ -\sin\alpha & \cos\alpha & 0 \\ 0 & 0 & 1 \end{pmatrix}, \quad R_x(\beta) = \begin{pmatrix} 1 & 0 & 0 \\ 0 & \cos\beta & \sin\beta \\ 0 & -\sin\beta & \cos\beta \end{pmatrix}, \quad R_z(\gamma) = \begin{pmatrix} \cos\gamma & \sin\gamma & 0 \\ -\sin\gamma & \cos\gamma & 0 \\ 0 & 0 & 1 \end{pmatrix}. Expanding this product yields the explicit elements of R: R = \begin{pmatrix} \cos\alpha \cos\gamma - \sin\alpha \sin\gamma \cos\beta & \cos\alpha \sin\gamma + \sin\alpha \cos\gamma \cos\beta & \sin\alpha \sin\beta \\ -\sin\alpha \cos\gamma - \cos\alpha \sin\gamma \cos\beta & -\sin\alpha \sin\gamma + \cos\alpha \cos\gamma \cos\beta & \cos\alpha \sin\beta \\ \sin\gamma \sin\beta & -\cos\gamma \sin\beta & \cos\beta \end{pmatrix}. This formulation follows the standard x-convention for proper Euler angles, as derived in classical mechanics texts. For Tait-Bryan angles, which use three distinct axes, the z-y-x sequence (often called yaw-pitch-roll, with angles ψ for yaw about z, θ for pitch about y, and φ for roll about x) produces R = R_z(\psi) R_y(\theta) R_x(\phi). The individual matrices are: R_z(\psi) = \begin{pmatrix} \cos\psi & \sin\psi & 0 \\ -\sin\psi & \cos\psi & 0 \\ 0 & 0 & 1 \end{pmatrix}, \quad R_y(\theta) = \begin{pmatrix} \cos\theta & 0 & -\sin\theta \\ 0 & 1 & 0 \\ \sin\theta & 0 & \cos\theta \end{pmatrix}, \quad R_x(\phi) = \begin{pmatrix} 1 & 0 & 0 \\ 0 & \cos\phi & \sin\phi \\ 0 & -\sin\phi & \cos\phi \end{pmatrix}. The expanded matrix elements are: R = \begin{pmatrix} \cos\psi \cos\theta & \sin\phi \sin\theta \cos\psi - \cos\phi \sin\psi & \cos\phi \sin\theta \cos\psi + \sin\phi \sin\psi \\ -\sin\psi \cos\theta & \cos\phi \cos\psi + \sin\phi \sin\theta \sin\psi & -\sin\phi \cos\psi + \cos\phi \sin\theta \sin\psi \\ \sin\theta & -\sin\phi \cos\theta & \cos\phi \cos\theta \end{pmatrix}. This z-y-x form is widely used in and applications. To convert from a rotation matrix back to Euler angles, the process involves extracting the angles using specific elements of R, typically via arctangent functions to handle the appropriate quadrants. For the z-x-z proper Euler convention, with R as above, the angle β is first obtained as \beta = \arccos(R_{33}), valid for \beta \in [0, \pi]. Then, \gamma = \atantwo(R_{31}, -R_{32}) and \alpha = \atantwo(R_{13}, R_{23}), ensuring the signs align with the convention. For the z-y-x Tait-Bryan convention, θ is extracted as \theta = \arcsin(R_{31}), valid for \theta \in [-\pi/2, \pi/2]. Subsequently, \psi = \atantwo(-R_{32}, R_{33}) and \phi = \atantwo(R_{21}, R_{11}). These formulas assume no singularities, where the pitch angle θ is zero or ±π/2, which would require special handling. Similar extractions apply to other sequences by selecting the appropriate matrix elements corresponding to the sine and differences of cosines. In all cases, the resulting rotation matrix R satisfies orthogonality (R^T R = I) and has determinant 1 (\det(R) = 1), confirming it represents a proper rotation in SO(3). These properties hold by construction from the individual rotation matrices, each of which is orthogonal with unit determinant.

To and from quaternions

Unit quaternions provide a compact and efficient representation for three-dimensional rotations, defined as q = w + x\mathbf{i} + y\mathbf{j} + z\mathbf{k} where w, x, y, z \in \mathbb{R} and \|q\| = \sqrt{w^2 + x^2 + y^2 + z^2} = 1. This normalization ensures that q and -q describe the same rotation, avoiding the gimbal lock inherent in Euler angle sequences. To convert Euler or Tait-Bryan angles to a unit quaternion, the rotations are composed using quaternion multiplication with half-angles, as a rotation by angle \alpha about an axis corresponds to the quaternion \cos(\alpha/2) + \sin(\alpha/2) \mathbf{u}, where \mathbf{u} is the unit axis vector. For the common ZYX Tait-Bryan sequence (yaw \psi, pitch \theta, roll \phi), the quaternion is obtained by q = q_z(\psi/2) q_y(\theta/2) q_x(\phi/2), where: q_x(\phi/2) = \cos(\phi/2) + \sin(\phi/2) \mathbf{i}, \quad q_y(\theta/2) = \cos(\theta/2) + \sin(\theta/2) \mathbf{j}, \quad q_z(\psi/2) = \cos(\psi/2) + \sin(\psi/2) \mathbf{k}. The components are then: \begin{align*} w &= \cos(\theta/2) \cos(\phi/2) \cos(\psi/2) + \sin(\theta/2) \sin(\phi/2) \sin(\psi/2), \\ x &= \cos(\theta/2) \sin(\phi/2) \cos(\psi/2) - \sin(\theta/2) \cos(\phi/2) \sin(\psi/2), \\ y &= \sin(\theta/2) \cos(\phi/2) \cos(\psi/2) + \cos(\theta/2) \sin(\phi/2) \sin(\psi/2), \\ z &= \cos(\theta/2) \cos(\phi/2) \sin(\psi/2) - \sin(\theta/2) \sin(\phi/2) \cos(\psi/2). \end{align*} This multiplication order reflects the intrinsic rotation sequence. Converting from a unit quaternion back to Euler or Tait-Bryan angles can be done directly using trigonometric identities, avoiding the need for an intermediate rotation matrix, though matrix methods remain common. For the ZYX sequence, the pitch angle \theta is extracted as \theta = \asin(2(w x - y z)), where q = (w, x, y, z), assuming the convention where positive pitch aligns with the y-axis. The yaw \psi and roll \phi follow from atan2 functions on combinations of the components, such as \psi = \atantwo(2(x y + w z), w^2 + x^2 - y^2 - z^2) and \phi = \atantwo(2(y z + w x), w^2 - x^2 - y^2 + z^2), with adjustments near singularities where |\theta| \approx \pm \pi/2. These direct formulas ensure numerical stability and efficiency, with errors on the order of machine precision in implementations. One key advantage of quaternions over Euler angles is their suitability for singularity-free interpolation, such as spherical linear interpolation (SLERP), which produces smooth, constant-speed paths between orientations without discontinuities. This makes quaternions particularly valuable in applications requiring animated or dynamic rotations, where Euler angles may introduce jumps or lose a degree of freedom at gimbal lock points.

Other representations

The axis-angle representation parameterizes a rotation by a unit axis vector \hat{n} and a rotation angle \theta about that axis, providing a compact alternative to Euler angles for certain computations. To extract the axis and angle from Euler angles, the equivalent rotation matrix is first computed from the sequence of Euler rotations, after which the inverse Rodrigues' rotation formula is applied: the angle \theta is obtained from \theta = \cos^{-1}\left(\frac{\operatorname{trace}(R) - 1}{2}\right), and the axis \hat{n} from the skew-symmetric components of the matrix logarithm \log(R), ensuring \theta \in [0, \pi]. Rodrigues parameters offer another three-parameter representation using a vector \mathbf{r} = \tan(\theta/2) \hat{n}, where \theta and \hat{n} are the axis-angle quantities; this form is derived similarly by converting Euler angles to the rotation matrix and solving for \mathbf{r} via the relation R = I + \frac{2}{1 + \mathbf{r} \cdot \mathbf{r}} ([\mathbf{r}]_\times + \mathbf{r} \mathbf{r}^T), providing a finite range without the discontinuities of Euler angles near \theta = \pi. The direction cosine matrix (DCM) is essentially synonymous with the rotation matrix, denoting the 3×3 orthogonal matrix whose elements are the cosines of the angles between the axes of two coordinate frames; in Euler angle contexts, it is constructed directly as the product of the individual rotation matrices for each angle in the chosen convention, such as R = R_z(\psi) R_y(\theta) R_x(\phi) for the ZYX sequence. Conversions from Euler angles to these representations are typically performed for improved numerical stability—avoiding gimbal lock singularities—or to interface with software that natively supports them, such as axis-angle formats in computer graphics engines.

Properties and Generalizations

Gimbal lock and singularities

Gimbal lock refers to the loss of one degree of freedom in the representation of three-dimensional rotations using Euler angles, occurring at specific configurations where the transformation relating angular rates to the angular velocity becomes singular. This phenomenon arises because Euler angles parameterize the special orthogonal group SO(3), a three-dimensional manifold, using three coordinates that fail to span the tangent space at certain points, resulting in a rank-deficient Jacobian matrix. For proper Euler angles, such as the z-x-z sequence, singularities occur when the intermediate angle θ equals 0 or π, where the first and third rotation axes align, making their contributions indistinguishable. In Tait-Bryan angles, like the z-y-x sequence commonly used in aerospace, the singularity happens when the intermediate pitch angle θ = ±π/2, again aligning axes and reducing the effective degrees of freedom from three to two. Mathematically, this singularity manifests in the kinematic relation between the angular velocity vector ω and the Euler angle rates \dot{\phi}, expressed as \boldsymbol{\omega} = \mathbf{T}(\boldsymbol{\phi}) \dot{\boldsymbol{\phi}}, where \mathbf{T} is a 3×3 transformation matrix dependent on the current angles \boldsymbol{\phi} = (\phi_1, \phi_2, \phi_3). At the singular configurations, \det(\mathbf{T}) = 0, causing the matrix to lose full rank and preventing unique recovery of the angle rates from the angular velocity, which complicates control and integration in dynamic systems. For example, in the z-y-x convention, \mathbf{T} includes terms like \cos\theta in the denominator, leading to singularity as \theta \to \pm \pi/2. From a geometric perspective, Euler angles provide local charts on the SO(3) manifold, but these charts exhibit polar-like singularities where the coordinate system degenerates, analogous to longitude undefined at the poles on a sphere. A single set of Euler angles covers SO(3) except at these singular points, requiring multiple charts—such as four Euler angle-based charts—to form a complete atlas without singularities across the entire manifold. To mitigate gimbal lock, one approach is sequence switching, where the Euler angle convention is dynamically changed near singularities to a different rotation order that avoids alignment, though this requires careful handling of continuity in applications like attitude determination. Another method involves using universal joints in physical mechanisms, which inherently limit motion ranges to prevent axis alignment and thus avoid lock, as seen in certain robotic wrists or vehicle drivetrains.

Geometric algebra formulation

In geometric algebra, rotations in three-dimensional Euclidean space are represented using rotors, which provide a coordinate-free framework for describing orientations. A rotor R for a rotation by an angle \theta in the plane defined by a unit bivector B (with B^2 = -1) is given by R = \exp\left( -\frac{B \theta}{2} \right), where the exponential is defined via its power series expansion. This formulation arises from the even subalgebra of the geometric algebra \mathcal{G}_3, where rotors form a group under multiplication and sandwich a vector \mathbf{v} to yield the rotated vector R \mathbf{v} \tilde{R}, with \tilde{R} the reverse of R. Euler angles can be expressed as a composition of three such rotors, corresponding to successive rotations about orthogonal axes. For a general Euler angle sequence, the total rotor is R = R_3 R_2 R_1, where each R_i = \exp\left( - \frac{\mathbf{e}_i \wedge \mathbf{e}_j \alpha_i}{2} \right) for angles \alpha_i and basis bivectors \mathbf{e}_i \wedge \mathbf{e}_j defining the planes of rotation (e.g., \mathbf{e}_1 \wedge \mathbf{e}_2 for rotation about \mathbf{e}_3). The order of multiplication reflects the sequence of rotations applied from right to left, enabling a direct algebraic representation without explicit matrix construction. This approach is detailed in the treatment of rotations in \mathcal{G}_3. The geometric algebra formulation offers advantages over matrix-based representations, including a unified treatment of rotations and Lorentz boosts through generalization to the spacetime algebra \mathcal{G}_{1,3}, where rotors encompass both via even elements. Composition of rotations is simplified as a straightforward product of rotors, preserving the geometric interpretation of planes and angles, which facilitates computations in physics and engineering without coordinate dependencies. As an example, the ZXZ Euler angle convention—common in quantum mechanics and rigid body dynamics—corresponds to rotations about the z-axis by \phi, then the new x-axis by \theta, and again the z-axis by \psi. In geometric algebra, this yields the rotor R = \exp\left( -\frac{\mathbf{e}_1 \wedge \mathbf{e}_2 \phi}{2} \right) \exp\left( -\frac{\mathbf{e}_2 \wedge \mathbf{e}_3 \theta}{2} \right) \exp\left( -\frac{\mathbf{e}_1 \wedge \mathbf{e}_2 \psi}{2} \right), where the bivectors \mathbf{e}_1 \wedge \mathbf{e}_2 and \mathbf{e}_2 \wedge \mathbf{e}_3 define the xy- and yz-planes, respectively. Applying this rotor to basis vectors reproduces the standard rotation matrix for the sequence.

Extensions to higher dimensions

The special orthogonal group SO(n), which describes rotations in n-dimensional Euclidean space, has dimension \frac{n(n-1)}{2}, requiring that many independent angles for a complete parameterization analogous to the three Euler angles in 3D. Generalizations of Euler angles to higher dimensions typically involve decomposing elements of SO(n) into a product of rotations in nested 2D planes, often using a chain of n-1 successive rotations about orthogonal axes or via maximal embedded subgroups, such as reducing to SO(n-1) iteratively. This approach extends the 3D concept of composing rotations around body-fixed or space-fixed axes but introduces more parameters, with explicit formulas for the resulting orthogonal matrices derived from angle compositions. In 4 dimensions, SO(4) requires 6 angles for full specification, and one common Euler-like parameterization leverages its isomorphism to the double cover SU(2) × SU(2)/ℤ₂, representing rotations as a pair of unit quaternions (each parameterized by 3 Euler angles) or as chained rotations in orthogonal planes, such as two independent 3D rotations combined via the group's structure. For instance, a general element can be expressed as a product of rotations in the (1,2), (1,3), (1,4), (2,3), (2,4), and (3,4) planes, though practical implementations often use recursive decompositions to avoid redundancy. These methods preserve some advantages of 3D Euler angles, like intuitive geometric interpretation, but the increased dimensionality amplifies singularities, where the parameterization fails or becomes non-unique, extending the gimbal lock issue to higher-order degeneracies. Higher-dimensional extensions face significant challenges, including a proliferation of singularities due to the topology of SO(n)—far more than the isolated gimbal locks in 3D—and elevated computational complexity from inverting the chained rotations to extract angles from a given matrix, often requiring iterative numerical methods. For n > 3, the Euler parameters (an overcomplete set satisfying a ) introduce additional mathematical singularities at certain points, such as when a base parameter vanishes, complicating applications in fields like attitude determination.

Applications

Aerospace and vehicle dynamics

In aerospace engineering, Euler angles, particularly the Tait-Bryan variant, are widely employed to represent the attitude of aircraft and spacecraft relative to an inertial frame. This sequence typically involves yaw (ψ) for heading, pitch (θ) for elevation, and roll (φ) for bank, allowing engineers to decompose complex three-dimensional orientations into intuitive sequential rotations about the vehicle's body axes. Such representations are essential in flight simulators and navigation systems, where they facilitate real-time computation of vehicle orientation for stability and control. Euler angles also play a key role in formulating the equations of motion for rigid body dynamics in vehicles. The angular velocity vector \boldsymbol{\omega} is expressed in terms of the time derivatives of these angles, enabling the application of Euler's rotational equations: \frac{d\mathbf{L}}{dt} + \boldsymbol{\omega} \times \mathbf{L} = \boldsymbol{\tau} where \mathbf{L} is the angular momentum, and \boldsymbol{\tau} is the applied torque. This framework is particularly valuable in aerospace for modeling the rotational dynamics of spacecraft, where attitude maneuvers must account for conservation of angular momentum under varying torques from thrusters or environmental disturbances. In spacecraft attitude control systems, parameterize the orientation for feedback loops that stabilize the vehicle during orbit adjustments or pointing tasks. For instance, control algorithms use these angles to command reaction wheels or thrusters, ensuring precise alignment despite singularities like near 90-degree pitch. Similarly, in launch vehicles such as rockets, systems for thrust vector control rely on Euler angle representations to adjust orientations, correcting deviations in real time. Software tools in extensively incorporate Euler angles for simulating . and Simulink's Aerospace Blockset provides the 6DOF (Euler Angles) block, which implements these representations in six-degrees-of-freedom models to predict translational and rotational responses under aerodynamic and gravitational forces. This enables engineers to design and validate control systems for applications ranging from autopilots to interplanetary probes.

Computer graphics and robotics

In , Euler angles are widely used to represent rotations of 3D objects and scenes in rendering pipelines. For instance, in , they facilitate the construction of rotation matrices by sequencing individual axis rotations, enabling efficient transformation of vertex coordinates during rendering. Similarly, game engines like employ Euler angles to define object orientations, allowing developers to specify rotations in intuitive terms such as yaw, pitch, and roll for character controllers and environmental elements. Tait-Bryan angles, a variant of Euler angles using rotations about distinct axes (typically Z-Y-X for yaw-pitch-roll), are particularly common for camera views, as they align with human perception of heading, elevation, and banking in virtual environments. In robotics, Euler angles play a key role in kinematic modeling, where forward kinematics computes end-effector pose from joint angles, and inverse kinematics solves for joint configurations to reach a target pose while respecting physical constraints. These angles parameterize rotational joints, enabling algorithms to enforce joint limits that prevent mechanical overextension or collisions in manipulators like industrial arms. For example, in six-degree-of-freedom robots, ZYZ Euler angles are often used to match the orientation of the end-effector to desired tasks, such as precise assembly, by decoupling translational and rotational components. A significant challenge with Euler angles in animations and robotic is , where alignment of rotation axes leads to loss of a degree of freedom, causing erratic or control issues during sequences like camera pans or arm swings. This is commonly mitigated by converting to quaternions for smoother, lock-free , as introduced in seminal work on quaternion-based animation curves. Such handling briefly references broader issues in Euler representations, ensuring robust path planning. As an example, Euler integration—linear interpolation between keyframe Euler angles—is applied in keyframe animation to generate intermediate rotations, providing predictable control over motion paths in tools like Unity's Animation window for character limb movements or object tumbling. This method ensures tangents align smoothly without the spherical twisting artifacts of quaternion slerp, though it requires careful axis ordering to avoid discontinuities.

Crystallography and materials science

In and , Euler angles provide a representation of crystal orientations, essential for quantifying preferred orientations or textures in polycrystalline materials such as metals and alloys. The Bunge convention, a specific proper Euler angle sequence denoted as (φ₁, Φ, φ₂), is the standard for describing the orientation of individual grains relative to the sample . In this convention, φ₁ represents the about the sample's (typically the z-axis), Φ is the tilt about the transverse (x-axis), and φ₂ is the about the 's (again z-axis). This triplet fully parameterizes the that aligns the with the macroscopic sample axes, enabling the analysis of during processes like rolling or annealing. The Bunge convention was formalized by Hans-Rudolf Bunge in the late to standardize measurements, particularly for preferred orientations in deformed metals, where traditional methods like required a consistent angular framework. Bunge's approach, introduced in his seminal 1969 monograph, facilitated the mathematical treatment of orientation distribution functions (ODFs), which statistically describe the volume fraction of crystals at each Euler angle triplet. This adoption in the marked a shift toward quantitative analysis, allowing researchers to correlate microstructural orientations with mechanical properties like in formability. Pole figures, which map the angular distribution of crystallographic poles (directions normal to lattice planes) onto a , rely on Euler angles to interpret patterns from techniques such as or . By converting measured pole densities into Euler angle space, researchers reconstruct full ODFs, revealing the three-dimensional that two-dimensional pole figures alone cannot capture. For instance, a strong {111} in rolled aluminum appears as clustered poles on a figure, corresponding to specific ranges in φ₁ and Φ. In (EBSD), Euler angles are directly computed from Kikuchi patterns to map local grain orientations in scanning electron microscopy, enabling three-dimensional reconstruction when combined with serial sectioning. EBSD software, such as those from EDAX/TSL or Oxford Instruments, outputs Bunge Euler angles for each pixel, facilitating of components like or Goss orientations in alloys. This application has become routine since the 1990s, supporting studies of recrystallization and phase transformations in materials.

Other fields

In astronomy, Euler angles are employed to describe the of satellites and planetary probes relative to inertial reference frames, facilitating precise orientation for instruments such as telescopes and sensors. This parameterization allows for the representation of rotations through sequences like roll-pitch-yaw, which are integrated into attitude determination algorithms using data from star trackers or gyroscopes. For instance, in missions like OrbView-2, Euler angles enable linearized models for design, though they are prone to singularities during large maneuvers. In , Euler angles play a key role in techniques for aligning multi-modal scans, such as MRI and volumes, to reconstruct three-dimensional patient anatomies and quantify joint motions. Automated 2D-3D registration algorithms use Euler angles (e.g., in the XZ'Y'' sequence) to parameterize rotations during optimization processes that match digitally reconstructed radiographs to fluoroscopic images, achieving root-mean-square errors of 0.6–2.2° for angular accuracy in motion analysis. This approach supports applications in orthopedics, where it enables non-invasive assessment of glenohumeral elevation, plane of elevation, and axial rotation without fiducial markers. In , Euler angles are utilized to specify the of symmetries in modeling seismic , which affects wave propagation velocities and in the Earth's subsurface. By representing the of elastic stiffness tensors through three angles (e.g., in the Bunge convention), researchers compute anisotropic seismic properties of polycrystalline aggregates, such as shales or rocks, using self-consistent approximation methods. Tools like AnisEulerSC integrate these angles to predict P- and S-wave anisotropies from microstructural data, aiding in the interpretation of coefficients and reflector / in . In , Euler angles parameterize rotations in the spin-1 representation of the SO(3) group, providing a basis for describing the orientation of vector operators like in systems with spin. For spin-1 particles, the rotation operator is expressed as a sequence of rotations about z, y, and z axes with angles α, β, γ, linking the special orthogonal group SO(3) to physical transformations while accounting for the double cover by SU(2) in half- cases. This formulation is essential for computing elements in and molecular spectra, where the angles facilitate the of rotation operators into irreducible representations.

References

  1. [1]
  2. [2]
    Formulae generales pro translatione quacunque corporum rigidorum
    Sep 25, 2018 · Formulae generales pro translatione quacunque corporum rigidorum ; English Title. General formulas for the translation of arbitrary rigid bodies ...
  3. [3]
    [PDF] Euler Angles, Unit Quaternions, and Rotation Vectors
    Oct 20, 2006 · Abstract. We present the three main mathematical constructs used to represent the attitude of a rigid body in three- dimensional space.Missing: authoritative | Show results with:authoritative
  4. [4]
    [PDF] MATH431: Gimbal Lock - UMD MATH
    Nov 3, 2021 · To achieve all possible orientations of an object in R3 there are three degrees of freedom that must be accounted for.
  5. [5]
    The Euler angle parameterization - Rotations
    In Euler's papers, he shows how three angles can be used to parameterize a rotation, and he also establishes expressions for the corotational components of the ...
  6. [6]
  7. [7]
  8. [8]
    [2006.00196] On the Work of Benjamin Olinde Rodrigues (1795-1851)
    May 30, 2020 · His chef-d'oeuvre is the work on Euclidean motion group in 1840. He invented Rodrigues expression of rotation and give explicit calculation ...
  9. [9]
    [PDF] ON QUATERNIONS By William Rowan Hamilton
    In the theory which Sir William Hamilton submitted to the Academy in November,. 1843, the name quaternion was employed to denote a certain quadrinomial ...
  10. [10]
  11. [11]
  12. [12]
    [1909.06669] Special Orthogonal Group SO(3), Euler Angles ... - arXiv
    Sep 14, 2019 · This paper gives an overview of the rotation matrix, attitude kinematics and parameterization. The four most frequently used methods of attitude ...
  13. [13]
    [PDF] Rotations and Orientation - UT Computer Science
    Turns out intrinsic rotation order is the reverse of extrinsic rotation order! e.g. intrinsic rotation (x y z) is equivalent to extrinsic rotation (z y x).
  14. [14]
    Proof of the extrinsic to intrinsic rotation transform
    Feb 7, 2015 · Any extrinsic rotation is equivalent to an intrinsic rotation by the same angles but with inverted order of elemental rotations, and vice-versa.Why does intrinsic euler angle rotation not equal extrinsic rotation?Extrinsic and intrinsic Euler angles to rotation matrix and backMore results from math.stackexchange.com
  15. [15]
    [PDF] Rotations in 3-Dimensional Space
    Because of this, the rotation group SO(3) is said to be non-Abelian. This means that the rotation group is noncommutative. Let us write R1 = R(n1,θ1) and R2 = R ...
  16. [16]
    [PDF] Computing Euler angles from a rotation matrix
    This document discusses a simple technique to find all possible Euler angles from a rotation matrix. Determination of Euler angles is sometimes a necessary step.
  17. [17]
    Linear Algebra, Part 1: Euler Theorem (Mathematica)
    Leonhard Euler proved the following theorem in 1775 that states that in three-dimensional space, any displacement of a rigid body such that a point on the ...
  18. [18]
    [PDF] Rotation Matrices and Quaternion - Lecture
    Euler Angles. • Proper Euler angles: z-x-z, x-y-x, y-z-y, z-y-z, x-z-x, y-x-y. • Tait-Bryan angles: x-y-z, y-z-x, z-x-y, x-z-y, z-y-x, y-x-z. 20. Page 21. z-y-x ...
  19. [19]
    Euler angles - Knowino
    May 10, 2013 · The angles are named after the 18th century mathematician Leonhard Euler who introduced in 1765 two of the three for an axially symmetric ...
  20. [20]
    Euler Angles | PhD - Guidance, Navigation and Control Engineering
    May 5, 2019 · Definition. Euler angles describe any arbitrary rotation using a set of 3 angles, each angle representing a single axis rotation that is ...<|control11|><|separator|>
  21. [21]
    Euler angle conventions - CCP-EM
    A common convention is to rotate α about the z-axis, β about the new y-axis, and γ about the final z-axis. These are rotations with respect to an internal ...
  22. [22]
  23. [23]
    Euler Angles | Academic Flight
    The idea of the Euler angles comes from the following. Take an aircraft in a default position such that its body frame coincides with the inertial frame,
  24. [24]
    Other Explicit Representation for the Orientation in Robotics: Euler ...
    Jul 28, 2021 · So, both sets of ZXZ Euler angles of (π,π/4,π/4), and (0,-π/4,5π/4) can represent the rotation matrix for the given orientation. Let's watch the ...Missing: uniqueness proper<|separator|>
  25. [25]
    [PDF] 3D Rigid Body Dynamics: Euler Angles - MIT OpenCourseWare
    Euler angles are particularly useful to describe the motion of a body that rotates about a fixed point, such. as a gyroscope or a top or a body that rotates ...Missing: zxz | Show results with:zxz
  26. [26]
    [PDF] Deriving Euler's Equation for Rigid-Body Rotation via Lagrangian ...
    Jun 16, 2023 · Comparing (13) with Euler's Equation (1) written in terms of the angular velocity implies ... (precession-nutation-spin) Euler angles, it follows ...
  27. [27]
    Precession and Forced Nutation of the Earth - Richard Fitzpatrick
    Unfortunately, the observed precession period of the Earth's axis of rotation about the normal to the ecliptic plane is approximately 25,800 years, so something ...
  28. [28]
    [PDF] 3D rotation sequences (eg, yaw/pitch/roll)
    Recall that the {3,2,1} Tait-Bryan rotation sequence is defined by three successive rotations. (yaw by , then pitch by , then roll by ), and thus may be ...
  29. [29]
    [PDF] Robot Dynamics Lecture Notes
    B.2 Rotation matrix to Euler angles. This section implements the extraction of Euler angles (ZYZ, ZXZ, ZYX and XYZ) from a given rotation matrix. 92. Page 98 ...<|control11|><|separator|>
  30. [30]
    [PDF] 10. Aircraft Equations of Motion - Translation, Rotation - 2018
    What are Euler angles? ❑ What rotation sequence is used to describe airplane attitude? ❑ What are properties of the rotation.
  31. [31]
  32. [32]
    How to avoid singularity when using Euler angles? - ResearchGate
    Aug 9, 2025 · In this paper, an algorithm is proposed to avoid singularity associated with the most famous minimum element attitude parametrization, Euler angle set.
  33. [33]
    Euler Angles -- from Wolfram MathWorld
    The three angles giving the three rotation matrices are called Euler angles. There are several conventions for Euler angles, depending on the axes about which ...Missing: authoritative sources
  34. [34]
    [PDF] Euler Angle Formulas - Geometric Tools
    Dec 1, 1999 · For example, one might want to factor a rotation as R = Rx(θx)Ry(θy)Rz(θz) for some angles θx, θy, and θz. The ordering is xyz. Five other ...
  35. [35]
    Rotation Matrix -- from Wolfram MathWorld
    When discussing a rotation, there are two possible conventions: rotation of the axes, and rotation of the object relative to fixed axes. ... v^'=R_thetav_0.
  36. [36]
  37. [37]
    Rotations, Orientation, and Quaternions - MATLAB & Simulink
    One advantage of quaternions over Euler angles is the lack of discontinuities. Euler angles have discontinuities that vary depending on the convention being ...
  38. [38]
    [PDF] MODERN ROBOTICS - Mech
    May 3, 2017 · ... Modern Robotics. Mechanics, Planning, and Control c Kevin M. Lynch and Frank C. Park. This preprint is being made available for personal use ...Missing: Tait- Bryan
  39. [39]
    Rotation Matrix (Direction Cosine Matrix (DCM)) - Academic Flight
    The rotation matrix B is often also called the Direction Cosine Matrix (DCM). We have seen in our linear algebra primer that the columns of the transformation ...
  40. [40]
    None
    ### Summary of Euler Angles and Related Concepts from the Document
  41. [41]
    [PDF] Universit¨at Stuttgart Fachbereich Mathematik
    We describe explicitly an atlas for the rotation group SO(3) consisting of four charts where each chart is defined by Euler angles or each chart is defined by ...
  42. [42]
    A minimal atlas for the rotation group SO(3)
    May 15, 2011 · We describe explicitly an atlas for the rotation group SO(3) consisting of four charts where each chart is defined by Euler angles or each chart is defined by ...Missing: poles | Show results with:poles
  43. [43]
    (PDF) How to Avoid Singularity for Euler Angle Set? - ResearchGate
    The proposed algorithm makes use of method of sequential rotation to avoid singularity associated with Eu-ler angle set. Further, a switching algorithm is also ...
  44. [44]
    [PDF] kinematics of hooke universal joint robot wrists
    angle of the driver shaft of a.Ilooke universal joint. 0,,, _/',t actuator angles which produce Euler angles of 0 and 0. 0,1, 0.q, 0€; joint angles of a ...
  45. [45]
    Geometric Algebra for Physicists - Chris Doran, Anthony Lasenby
    This book is a complete guide to the current state of the subject with early chapters providing a self-contained introduction to geometric algebra. Topics ...
  46. [46]
    [PDF] Some notes on Euler Angles. - Peeter Joot's Blog
    In [Doran and Lasenby(2003)] section 2.7.5 the Euler angle formula is devel- oped for {z, x0, z00} axis rotations by {φ, θ, ψ} respectively.
  47. [47]
    Rotations in classical mechanics using geometric algebra - arXiv
    Oct 30, 2022 · In geometric algebra, the rotation of a vector is described using rotors. Rotors are phasors where the imaginary number has been replaced by a oriented plane ...
  48. [48]
    Euler angle in nLab
    Jun 17, 2021 · The 3 classical Euler angles furnish a global parametrization and a smooth coordinate system on a big dense cell of the Lie group SO(3) of ...
  49. [49]
  50. [50]
    [PDF] Principal Rotation Representations of Proper NxN Orthogonal Matrices
    Classical Rodrigues Parameters: This three parameter set, also referred to as the Gibbs vec- tor, is proportional to Euler's principal rotation vector. The ...
  51. [51]
  52. [52]
    [PDF] Navigation Filter Best Practices
    ... attitude estimation used a roll, pitch, yaw sequence of Euler or Tait-Bryan angles discussed in Section 9.7 [21,22]. This is a very useful representation if ...
  53. [53]
    Axes Transformations - Aircraft Flight Mechanics by Harry Smith, PhD
    Aircraft axes transformation uses Euler angles (yaw, pitch, roll) and a transformation matrix to define coordinates between Earth and body axes.
  54. [54]
    [PDF] Euler's Equations - 3D Rigid Body Dynamics - MIT OpenCourseWare
    We now turn to the task of deriving the general equations of motion for a three-dimensional rigid body. These equations are referred to as Euler's equations ...Missing: aerospace | Show results with:aerospace
  55. [55]
    Euler Angles | SpringerLink
    Cite this chapter. Markley, F.L., Crassidis, J.L. (2014). Euler Angles. In: Fundamentals of Spacecraft Attitude Determination and Control. Space Technology ...
  56. [56]
    [PDF] l&XJ.J~E WI - ATTITUDE CONTROL IN SPACE '.
    The Euler angles (#, 6, Q) are with respect to a specified reference frame ... A gimbal coordinate system (TG, TG, kc) is rigidly attached to the ...
  57. [57]
    6DOF (Euler Angles) - Equations of Motion - MathWorks
    The 6DOF (Euler Angles) block implements the Euler angle representation of six-degrees-of-freedom equations of motion.
  58. [58]
    [PDF] OpenGL FAQ and Troubleshooting Guide
    ... Euler angles. Delphi code for performing basic vector, matrix ... Also most basic 3D computer graphics books cover it, because it's not. OpenGL−specific.
  59. [59]
    Editing glTF: Possibilities, Challenges, and Insights from UX3D
    Aug 4, 2020 · For example, DCC tools often use human-friendly Euler angles when constructing a 3D model and retain information about the order in which the ...
  60. [60]
    Chapter 6. Inverse kinematics
    Since the only degrees of freedom that affect orientation are q4, q5, and q6, we can solve for the ZYZ euler angles to achieve RD. First, let us derive the ...
  61. [61]
    Inverse Kinematics – Modeling, Motion Planning, and Control of ...
    To solve the inverse orientation problem, we take the help of Euler angles. Let us first take a brief look at Euler angles and how does it help us in solving ...
  62. [62]
    [PDF] Animating Rotation with Quaternion Curves
    Translations combine by adding vectors; rotations, by multiplying quaternions. The separate axes translations don't interact; the axes of rotations must.
  63. [63]
    Rotation in animations - Unity - Manual
    Oct 7, 2025 · Euler Angles interpolation applies the full range of motion to the GameObject specified by the angles you enter; if the rotation is greater than ...
  64. [64]
    Tutorial: Crystal orientations and EBSD — Or which way is up?
    It is important to note that the Euler angles that appear in the orientation description in the Bunge convention, denote such an “active” rotation of the ...<|control11|><|separator|>
  65. [65]
    None
    Below is a merged summary of Euler Angle Models in Spacecraft Attitude Determination and Control, consolidating all information from the provided segments into a comprehensive response. To retain maximum detail and ensure clarity, I will use a table in CSV format to organize the key aspects (Role, Advantages, Issues, and Useful URLs) while incorporating additional details from the text into the descriptions. The table will be followed by a narrative summary to tie together any remaining context not easily captured in the table.
  66. [66]
    Euler Angles - Satellite Wiki
    Euler angles captures this idea. This is a set of three 'ordered' rotation parameters. So, each parameter corresponds to rotation about a specific axis.
  67. [67]
    MRI VS CT-BASED 2D-3D AUTO-REGISTRATION ACCURACY ...
    We developed an automatic 2D-3D registration program that works with both CT- and MRI-based image volumes for quantifying joint motions.
  68. [68]
    [PDF] Three-Dimensional Object Registration Using ... - DSpace@MIT
    are the pitch, roll and yaw Euler angles, 0, 5 and 0, required for registration as ... types of scans, such as MRI and CT data for medical applications or ...
  69. [69]
    Seismic anisotropy — Introduction | GEOPHYSICS - SEG Library
    Jun 19, 2017 · Their results are expressed in terms of six P-wave “weak-anisotropy” parameters, three Euler angles specifying the symmetry-plane orientation, ...
  70. [70]
    AnisEulerSC: A MATLAB program combined with MTEX for ...
    Here, we present a MATLAB-based software, AnisEulerSC (Anisotropy from Euler angles using Self-Consistent approximation), for modeling the anisotropic seismic ...
  71. [71]
    [PDF] Rotations in Quantum Mechanics, and Rotations of Spin-1
    As in the notation SO(3), the S stands for “special,” which means the determinant is +1. The significance of SU(2) in the present discussion is that every ...