Fact-checked by Grok 2 weeks ago

Cylindrical coordinate system

The cylindrical coordinate system is a three-dimensional curvilinear coordinate system that extends the two-dimensional polar coordinate system by adding a vertical height coordinate, allowing for the specification of points in space using radial distance, angular position, and elevation. In this system, each point is represented by the triplet (r, \theta, z), where r \geq 0 is the radial distance from the z-axis in the xy-plane, \theta is the azimuthal angle measured counterclockwise from the positive x-axis (typically ranging from 0 to $2\pi), and z is the height along the z-axis, identical to the Cartesian z-coordinate. The conversion between cylindrical and Cartesian coordinates is straightforward and bidirectional: the Cartesian coordinates are obtained via x = r \cos \theta, y = r \sin \theta, and z = z, while the inverse yields r = \sqrt{x^2 + y^2}, \theta = \tan^{-1}(y/x) (with appropriate quadrant adjustment), and z = z. This system is orthogonal, with scale factors of 1 for r and z, and r for \theta, which facilitates the computation of differentials, gradients, and integrals in . Cylindrical coordinates are particularly valuable in applications involving , such as modeling cylindrical pipes, analyzing electromagnetic fields around wires, or solving partial differential equations like the in regions with rotational invariance around an . They simplify the of surfaces like cylinders (r = constant), cones (z = k r), and paraboloids (z = r^2), making them essential in , physics, and engineering for efficient problem-solving in symmetric geometries.

Definition and Conventions

Basic Components

The cylindrical coordinate system extends the two-dimensional into three dimensions by incorporating a vertical , providing a natural framework for problems exhibiting around a central . It specifies the position of a point in using three orthogonal coordinates: the radial distance \rho (sometimes denoted as r) measured from the z- in the xy-plane, the azimuthal \phi measured from the positive x- in the xy-plane, and the z along the z-. These coordinates are defined such that \rho \geq 0, \phi \in [0, 2\pi) or equivalently \phi \in (-\pi, \pi], and z \in \mathbb{R}. Geometrically, a point with coordinates (\rho, \phi, z) is located by first projecting onto the xy-plane at a distance \rho from the along the direction specified by \phi, then moving to the z-axis by distance z. This system aligns with the geometry of cylindrical objects, where surfaces of constant \rho form infinite cylinders coaxial with the z-axis, surfaces of constant \phi define half-planes containing the z-axis, and surfaces of constant z are horizontal planes perpendicular to the z-axis. The orthogonal nature of these coordinate surfaces facilitates analysis in contexts like and involving . For visualization, consider the of the , represented as (\rho = 0, \phi = 0, z = 0), which coincides with the Cartesian . Points along the z-axis, such as (0, 0, 5), have \rho = 0 regardless of \phi, emphasizing the axis as a degenerate of zero . This setup allows straightforward plotting of structures like or wells by varying \rho while keeping \phi and z fixed along generatrices.

Uniqueness and Representations

In cylindrical coordinates, a point in three-dimensional space is specified by the triple (\rho, \phi, z), where the azimuthal angle \phi is periodic with period $2\pi. Consequently, the representations (\rho, \phi, z) and (\rho, \phi + 2\pi k, z) for any integer k describe the identical point, leading to infinitely many equivalent coordinate triples for any given location off the z-axis./08:_Some_Curvilinear_Coordinate_Systems/8.02:_Spherical_and_Cylindrical_Coordinates) This non-uniqueness is particularly pronounced for points on the z-axis, where \rho = 0; here, the angle \phi becomes arbitrary and undefined in a directional sense, allowing any value of \phi to pair with \rho = 0 and a fixed z to represent the same axial point./08:_Some_Curvilinear_Coordinate_Systems/8.02:_Spherical_and_Cylindrical_Coordinates) For instance, the point at Cartesian coordinates (0, 0, 1) corresponds to (0, \phi, 1) for arbitrary \phi. To establish a unique representation, standard conventions restrict \rho \geq 0 and confine \phi to a principal , such as [0, 2\pi) or (-\pi, \pi], while treating points with \rho = 0 separately by assigning a conventional \phi value (often 0) or acknowledging their arbitrariness. An illustrative example is the point at Cartesian coordinates (1, 0, 0), which can be represented as (\rho = 1, \phi = 0, z = 0) or (\rho = 1, \phi = 2\pi, z = 0), but the restriction \phi \in [0, 2\pi) selects the former as principal. In rare specialized contexts, such as certain analyses in particle trajectometry, a signed \rho (negative radius) may be employed to extend representations, where negative \rho effectively corresponds to a positive with \phi shifted by \pi, though this is non-standard and typically avoided in general usage.

Standard Conventions

In the cylindrical coordinate system, notation for the coordinates varies across disciplines and standards. The (ISO 31-11) recommends using \rho for the radial from the z-, \phi for the azimuthal , and z for the height along the . However, many mathematical texts and physics applications commonly employ r for the radial and \theta for the azimuthal , paired with z./12%3A_Vectors_in_Space/12.07%3A_Cylindrical_and_Spherical_Coordinates) This preference for \phi in mathematics and \theta in physics stems from conventions in related systems like spherical coordinates, where \theta often denotes the polar in physics contexts. The orientation of the azimuthal angle follows a right-handed , with \phi or \theta measured counterclockwise from the positive x-axis in the xy-plane. The z-axis is typically directed upward, aligning with the standard orientation of the Cartesian z-axis./12%3A_Vectors_in_Space/12.07%3A_Cylindrical_and_Spherical_Coordinates) Standard ranges for the coordinates ensure unique representations, with \rho \geq 0, \phi \in [0, 2\pi) or alternatively \phi \in (-\pi, \pi], and z \in (-\infty, \infty). The choice of interval for the azimuthal angle accommodates periodicity while avoiding redundancy, though points on the z-axis exhibit non-uniqueness due to the angle's indeterminacy. Cylindrical coordinates originated in the as part of developments in , with Leonhard Euler introducing their explicit use in three dimensions around the 1760s for solving problems in . Standardization occurred in the through texts, such as those by , which formalized their role in multivariable analysis. Unlike two-dimensional polar coordinates, which specify position with only radial distance and azimuthal angle, cylindrical coordinates extend this framework to three dimensions by incorporating the z-coordinate, thereby inheriting polar conventions for the radial and angular components while adding axial height.

Coordinate Transformations

From Cartesian Coordinates

The transformation from Cartesian coordinates (x, y, z) to cylindrical coordinates (r, \theta, z) involves expressing the radial distance r from the z-axis, the azimuthal angle \theta in the xy-plane, and the height z. The radial coordinate r is the perpendicular distance from the point to the z-axis, which forms the of a with legs of lengths |x| and |y| in the xy-plane; by the , r = \sqrt{x^2 + y^2}. The azimuthal \theta measures the counterclockwise from the positive x-axis to the line connecting the origin to the projection of the point in the xy-plane; it is computed as \theta = \atantwo(y, x), where the two-argument arctangent function ensures the correct and provides a full of -\pi < \theta \leq \pi (or equivalently $0 \leq \theta < 2\pi by adjustment if needed). The z-coordinate remains unchanged, as z = z, preserving the along the . Special cases arise when the point lies on the z-axis, where x = 0 and y = 0, yielding r = 0; in this situation, \theta is undefined, as the angle lacks a unique direction from the origin in the xy-plane, though it may be assigned arbitrarily for continuity in some applications. For example, consider the point (3, 4, 5) in Cartesian coordinates. Here, r = \sqrt{3^2 + 4^2} = 5, \theta = \atantwo(4, 3) \approx 0.927 radians (or about 53.13 degrees), and z = 5.

To Cartesian Coordinates

The transformation from cylindrical coordinates (r, \theta, z) to Cartesian coordinates (x, y, z) maps a point in by projecting its radial distance and azimuthal angle onto the xy-plane while preserving the vertical coordinate. This conversion leverages the relationship between cylindrical and polar coordinates in the plane perpendicular to the z-axis. The explicit formulas for the transformation are given by \begin{align*} x &= r \cos \theta, \\ y &= r \sin \theta, \\ z &= z. \end{align*} These equations arise from the geometry of the xy-plane, where the point (r, \theta) corresponds to a position on a circle of radius r centered at the origin. Specifically, projecting the cylindrical point onto this plane forms a right triangle with hypotenuse r, adjacent side x along the x-axis, and opposite side y along the y-axis; the angle \theta is measured from the positive x-axis. Applying the trigonometric definitions for cosine and sine on the unit circle—extended to radius r—yields x = r \cos \theta and y = r \sin \theta, with the z-coordinate remaining unchanged due to the cylindrical system's axial symmetry. To verify the transformation, substitute the formulas back into the inverse relations, confirming consistency. This yields r = \sqrt{x^2 + y^2} = \sqrt{(r \cos \theta)^2 + (r \sin \theta)^2} = \sqrt{r^2 (\cos^2 \theta + \sin^2 \theta)} = \sqrt{r^2 \cdot 1} = r, using the Pythagorean identity \cos^2 \theta + \sin^2 \theta = 1. Similarly, \tan \theta = y/x = (r \sin \theta)/(r \cos \theta) = \tan \theta, which holds provided \cos \theta \neq 0, with adjustments for quadrants based on the signs of x and y. For example, consider the point (r = 5, \theta = \pi/4, z = 0) in cylindrical coordinates. Applying the transformation gives x = 5 \cos(\pi/4) = 5 \cdot (\sqrt{2}/2) \approx 3.536, y = 5 \sin(\pi/4) \approx 3.536, and z = 0, corresponding to a point on the xy-plane at 45 degrees from the x-axis.

Relation to Spherical Coordinates

The cylindrical and spherical coordinate systems share the azimuthal angle, which measures the rotation around the z-axis from the positive x-axis in the xy-plane. In the spherical system, points are defined by the radial distance \rho from the origin (\rho \geq 0), the polar angle \theta from the positive z-axis ($0 \leq \theta \leq \pi), and the azimuthal angle \phi ($0 \leq \phi < 2\pi). The transformation from spherical coordinates (\rho, \theta, \phi) to cylindrical coordinates (r, \theta_\text{cyl}, z) is given by r = \rho \sin \theta, \quad z = \rho \cos \theta, \quad \theta_\text{cyl} = \phi, where r is the perpendicular distance from the z-axis. The inverse transformation from cylindrical to spherical coordinates is \rho = \sqrt{r^2 + z^2}, \quad \theta = \atan2(r, z), \quad \phi = \theta_\text{cyl}, with \theta computed using the two-argument arctangent function to ensure the correct quadrant. Geometrically, the cylindrical system projects the spherical coordinates onto an infinite cylinder aligned with the z-axis, where r captures the horizontal distance from this axis and z the vertical position, while the polar angle \theta determines the relative contributions of r and z to the overall radial distance \rho. For example, the point (\rho = 5, \theta = \pi/3, \phi = 0) in spherical coordinates corresponds to r = 5 \sin(\pi/3) = 5\sqrt{3}/2 \approx 4.33, z = 5 \cos(\pi/3) = 2.5, and \theta_\text{cyl} = 0 in cylindrical coordinates.

Differential Elements

Line Element

In cylindrical coordinates, the infinitesimal line element ds, which represents the arc length along a curve, is given by the metric ds^2 = dr^2 + r^2 d\theta^2 + dz^2, where r is the radial distance, \theta is the azimuthal angle, and z is the axial coordinate. This expression is derived by considering the position vector in cylindrical coordinates and its differential d\mathbf{r} = dr \, \hat{r} + r \, d\theta \, \hat{\theta} + dz \, \hat{z}, where \hat{r}, \hat{\theta}, and \hat{z} are the orthogonal unit vectors. Since the coordinate directions are mutually perpendicular, the squared magnitude follows from the dot product ds^2 = d\mathbf{r} \cdot d\mathbf{r}, yielding the sum of the squared infinitesimal displacements in each direction via the Pythagorean theorem applied to orthogonal components. The scale factors for this orthogonal curvilinear system are h_r = 1, h_\theta = r, and h_z = 1, such that the general form ds^2 = h_r^2 dr^2 + h_\theta^2 d\theta^2 + h_z^2 dz^2 reproduces the metric. The terms in the line element have clear geometric interpretations: dr contributes the radial displacement directly, r \, d\theta accounts for the arc length in the azimuthal direction (as the infinitesimal angle d\theta subtends an arc of length proportional to the radius r), and dz provides the vertical displacement along the axis. These scale factors arise from the coordinate transformation from Cartesian coordinates, where the azimuthal component stretches by the factor r due to the circular geometry. A practical application of the is calculating the of a circle at fixed r and z, where dr = 0 and dz = 0, so ds^2 = r^2 d\theta^2 and ds = r \, d\theta. Integrating over one full from \theta = 0 to $2\pi gives the \int_0^{2\pi} r \, d\theta = 2\pi r.

In cylindrical coordinates (r, \theta, z), the infinitesimal dV used in triple integrals is given by dV = r \, dr \, d\theta \, dz. This form arises from the in multiple integrals, where the volume scales by the absolute value of the determinant of the transformation from Cartesian coordinates. To derive this, consider the coordinate transformation x = r \cos \theta, y = r \sin \theta, z = z. The matrix of partial derivatives is \begin{vmatrix} \cos \theta & -r \sin \theta & 0 \\ \sin \theta & r \cos \theta & 0 \\ 0 & 0 & 1 \end{vmatrix}, with determinant r. Thus, dV = |\det J| \, dx \, dy \, dz = r \, dr \, d\theta \, dz. Geometrically, this represents a small cylindrical shell segment: the cross-sectional area in the r-\theta plane is the polar area element r \, dr \, d\theta, extruded along the z-direction by dz. For example, the volume of a right circular cylinder of radius R and height H is computed as V = \int_0^H \int_0^{2\pi} \int_0^R r \, dr \, d\theta \, dz = \pi R^2 H.

Surface Elements

In cylindrical coordinates (r, \theta, z), the principal coordinate surfaces consist of cylinders at constant r, half-planes at constant \theta, and horizontal planes at constant z. The differential surface element d\mathbf{S} on each surface is obtained by taking the magnitude of the cross product of the partial derivatives of the position vector with respect to the two varying coordinates, which yields the area scalar dS and the unit normal vector pointing in the direction of the cross product. For a surface of constant r (a cylindrical surface), parametrize the position vector as \vec{r}(\theta, z) = r \cos \theta \, \hat{i} + r \sin \theta \, \hat{j} + z \, \hat{k}. The partial derivatives are \partial \vec{r}/\partial \theta = (-r \sin \theta \, \hat{i} + r \cos \theta \, \hat{j}) and \partial \vec{r}/\partial z = \hat{k}. Their cross product is r (\cos \theta \, \hat{i} + \sin \theta \, \hat{j}) = r \hat{r}, so the magnitude gives the scalar surface element dS = r \, d\theta \, dz with outward normal \hat{r} for a closed cylindrical volume. For a surface of constant \theta (a radial half-plane), parametrize as \vec{r}(r, z) = r \cos \theta \, \hat{i} + r \sin \theta \, \hat{j} + z \, \hat{k}. The partial derivatives are \partial \vec{r}/\partial r = \cos \theta \, \hat{i} + \sin \theta \, \hat{j} = \hat{r} and \partial \vec{r}/\partial z = \hat{k}. Their cross product is \hat{r} \times \hat{k} = -\hat{\theta}, yielding dS = dr \, dz with normal -\hat{\theta}. For a surface of constant z (a horizontal disk), parametrize as \vec{r}(r, \theta) = r \cos \theta \, \hat{i} + r \sin \theta \, \hat{j} + z \, \hat{k}. The partial derivatives are \partial \vec{r}/\partial r = \hat{r} and \partial \vec{r}/\partial \theta = r \hat{\theta}. Their cross product is \hat{r} \times (r \hat{\theta}) = r \hat{z}, so the magnitude is r and dS = r \, dr \, d\theta with normal \hat{z} (or -\hat{z} depending on the order of parameters, outward for the top of a closed volume). These elements can be derived alternatively using the scale factors of the cylindrical metric (h_r = 1, h_\theta = r, h_z = 1), where the surface element on the face orthogonal to the i-th coordinate is dS = h_j h_k \, du_j \, du_k. For instance, the lateral surface area of a finite cylinder of radius r and height H is computed as \int_0^H \int_0^{2\pi} r \, d\theta \, dz = 2\pi r H.

Vector Calculus Operations

Gradient and Divergence

In cylindrical coordinates (r, \theta, z), the gradient of a scalar function f(r, \theta, z) is given by \nabla f = \frac{\partial f}{\partial r} \hat{e}_r + \frac{1}{r} \frac{\partial f}{\partial \theta} \hat{e}_\theta + \frac{\partial f}{\partial z} \hat{e}_z, where \hat{e}_r, \hat{e}_\theta, and \hat{e}_z are the orthonormal unit vectors in the radial, azimuthal, and axial directions, respectively. This expression arises from the general form of the in orthogonal (u_1, u_2, u_3), defined as \nabla f = \sum_{i=1}^3 \frac{1}{h_i} \frac{\partial f}{\partial u_i} \hat{e}_i, where h_i are the scale factors associated with each coordinate direction. For cylindrical coordinates, the scale factors are h_r = 1, h_\theta = r, and h_z = 1, derived from the ds^2 = dr^2 + r^2 d\theta^2 + dz^2. Substituting these scale factors yields the specific formula above. The divergence of a vector field \mathbf{A} = A_r \hat{e}_r + A_\theta \hat{e}_\theta + A_z \hat{e}_z in cylindrical coordinates is \nabla \cdot \mathbf{A} = \frac{1}{r} \frac{\partial (r A_r)}{\partial r} + \frac{1}{r} \frac{\partial A_\theta}{\partial \theta} + \frac{\partial A_z}{\partial z}. This follows the general divergence formula in orthogonal , \nabla \cdot \mathbf{A} = \frac{1}{h_1 h_2 h_3} \left[ \frac{\partial (h_2 h_3 A_1)}{\partial u_1} + \frac{\partial (h_1 h_3 A_2)}{\partial u_2} + \frac{\partial (h_1 h_2 A_3)}{\partial u_3} \right], again using the cylindrical scale factors h_r = 1, h_\theta = r, h_z = 1. The derivation leverages the infinitesimal dV = h_1 h_2 h_3 \, du_1 du_2 du_3 = r \, dr \, d\theta \, dz and applies the definition of as the limit of through a small . As an illustrative example, consider a radial flow vector field \mathbf{A} = A_r(r) \hat{e}_r with no azimuthal or axial components. The divergence simplifies to \nabla \cdot \mathbf{A} = \frac{1}{r} \frac{d (r A_r)}{d r}, which equals zero if r A_r is constant, corresponding to incompressible radial flow.

Curl and Laplacian

In cylindrical coordinates (r, \theta, z), the curl of a vector field \mathbf{A} = A_r \hat{e}_r + A_\theta \hat{e}_\theta + A_z \hat{e}_z is given by \nabla \times \mathbf{A} = \left( \frac{1}{r} \frac{\partial A_z}{\partial \theta} - \frac{\partial A_\theta}{\partial z} \right) \hat{e}_r + \left( \frac{\partial A_r}{\partial z} - \frac{\partial A_z}{\partial r} \right) \hat{e}_\theta + \frac{1}{r} \left( \frac{\partial (r A_\theta)}{\partial r} - \frac{\partial A_r}{\partial \theta} \right) \hat{e}_z. This expression arises from the general formula for the curl in orthogonal curvilinear coordinates, which employs a determinant-like structure incorporating the scale factors h_r = 1, h_\theta = r, and h_z = 1 to account for the varying metric along the azimuthal direction. The Laplacian of a scalar f(r, \theta, z) in cylindrical coordinates is \nabla^2 f = \frac{1}{r} \frac{\partial}{\partial r} \left( r \frac{\partial f}{\partial r} \right) + \frac{1}{r^2} \frac{\partial^2 f}{\partial \theta^2} + \frac{\partial^2 f}{\partial z^2}. This operator is derived as the of the , \nabla^2 f = \nabla \cdot (\nabla f), substituting the cylindrical forms of these operators while applying the appropriate scale factors. For problems exhibiting where f is independent of z, the Laplacian simplifies to the two-dimensional polar form \frac{1}{r} \frac{\partial}{\partial r} \left( r \frac{\partial f}{\partial r} \right) + \frac{1}{r^2} \frac{\partial^2 f}{\partial \theta^2}, which is particularly useful in analyzing radially symmetric phenomena such as heat diffusion in disks.

Applications

Cylindrical harmonics arise as the complete set of solutions to , ∇²ψ = 0, in cylindrical coordinates (ρ, φ, z), obtained through the method of ./02%3A_Charges_and_Conductors/2.10%3A_Variable_Separation__Cylindrical_Coordinates) Assuming a product solution of the form ψ(ρ, φ, z) = R(ρ) Φ(φ) Z(z), substitution into yields three independent ordinary differential equations for the radial, azimuthal, and axial functions. The azimuthal equation is d²Φ/dφ² + m² Φ = 0, where m is an to ensure single-valuedness of ψ under φ → φ + 2π, with solutions Φ(φ) = e^{imφ} (or equivalently and cosines for real forms). The axial equation takes the form d²Z/dz² - k² Z = 0 for bounded or decaying solutions along z, yielding Z(z) = e^{±kz}; alternatively, for finite domains, sinh(kz) or (kz) are used to satisfy conditions. The radial equation becomes the ρ² d²R/dρ² + ρ dR/dρ + (k² ρ² - m²) R = 0 for oscillatory behavior, with solutions R(ρ) = J_m(kρ) and Y_m(kρ), the of the first and second kind; for evanescent cases (separation constant +k²), modified Bessel functions I_m(kρ) and K_m(kρ) apply. The general cylindrical harmonic is thus ψ_{mkn}(ρ, φ, z) = [A J_m(k_n ρ) + B Y_m(k_n ρ)] e^{imφ} e^{±k_n z}, where k_n are determined by conditions (e.g., zeros of J_m for Dirichlet conditions on ρ = a), with m and n indexing the modes; superpositions over m, n sum to the full solution. For infinite or semi-infinite cylinders, transforms in z replace the discrete exponentials. These harmonics find applications in for potential distributions in cylindrical domains, such as charged infinite cylinders or cables, and in for describing modes where the fields satisfy similar Helmholtz equations derived from Laplace's for time- cases.

Kinematics and Dynamics

In cylindrical coordinates, the of a particle describe its , , and using the radial r, azimuthal \theta, and axial coordinate z. The of a particle is given by \vec{\rho} = r \hat{e}_r + z \hat{e}_z, where \hat{e}_r and \hat{e}_z are the unit vectors in the radial and axial directions, respectively. The accounts for the time derivatives of these coordinates and the rotation of the basis vectors, yielding \vec{v} = \dot{r} \hat{e}_r + r \dot{\theta} \hat{e}_\theta + \dot{z} \hat{e}_z, where \hat{e}_\theta is the azimuthal unit vector and the term r \dot{\theta} represents the tangential speed due to angular motion. This expression arises from differentiating the with respect to time, incorporating the perpendicularity of the unit vectors \hat{e}_r and \hat{e}_\theta. For acceleration, further differentiation introduces centripetal, Coriolis, and tangential components: \vec{a} = (\ddot{r} - r \dot{\theta}^2) \hat{e}_r + (r \ddot{\theta} + 2 \dot{r} \dot{\theta}) \hat{e}_\theta + \ddot{z} \hat{e}_z. The radial component includes the centripetal acceleration -r \dot{\theta}^2 directed inward, while the azimuthal component features the Coriolis term $2 \dot{r} \dot{\theta} that couples radial and angular motions. These forms are derived by applying the chain rule to the velocity components, considering the time-varying nature of the unit vectors. In dynamics, Newton's second law is applied component-wise in cylindrical coordinates to relate forces to accelerations. For a particle of mass m, the equations of motion are m (\ddot{r} - r \dot{\theta}^2) = F_r, m (r \ddot{\theta} + 2 \dot{r} \dot{\theta}) = F_\theta, and m \ddot{z} = F_z, where F_r, F_\theta, and F_z are the force components. These decoupled equations facilitate analysis of problems involving rotational symmetry, such as motion in gravitational fields or constrained systems like beads on rotating wires.

References

  1. [1]
    Calculus III - Cylindrical Coordinates - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will define the cylindrical coordinate system, an alternate coordinate system for the three dimensional coordinate system ...
  2. [2]
    None
    ### Summary of Cylindrical Coordinate System
  3. [3]
    Cylindrical coordinates - Dynamics
    Cylindrical coordinates are defined with respect to a set of Cartesian coordinates, and can be converted to and from these coordinates using the atan2 function ...
  4. [4]
    [PDF] a. cylindrical and spherical coordinate systems
    defined by r = constant. Thus the three orthogonal surfaces defining the. cylindrical coordinates of a point are r = constant, ø constant, and z = constant. ...
  5. [5]
    [PDF] 8.01SC S22 Chapter 3: Vectors - MIT OpenCourseWare
    This symmetry property of cylinders suggests a coordinate system, called a cylindrical coordinate system, that makes the symmetrical property under rotations.
  6. [6]
    17.6 Cylindrical and Spherical Coordinates
    The cylindrical coordinate system is the simplest, since it is just the polar coordinate system plus a z coordinate. A typical small unit of volume is the ...Missing: components | Show results with:components
  7. [7]
    Cylindrical coordinates - Math Insight
    Just as with polar coordinates, we usually limit 0≤θ<2π and r≥0 to descrease the non-uniqueness of cylindrical coordinates. However, when r=0, there is a non- ...Missing: periodicity | Show results with:periodicity
  8. [8]
    [PDF] Statistics for trajectometry - EPJ Web of Conferences
    ... cylindrical coordinates (r, Φ, z) are natural ... cylindrical coordinates of the perigee ... In this example, the signed radius R is negative (ϕ decreases with ...
  9. [9]
    A.7 ISO Coordinate System Notation
    In the ISO convention the symbols , ρ , ϕ and z are used (instead of , r , θ and z ) for cylindrical coordinates. distance from ...<|separator|>
  10. [10]
    [PDF] CONVENTIONS FOR SPHERICAL COORDINATES
    Oct 18, 2002 · American texts use θ for equatorial, φ for zenith, and ρ for radial. Other use θ for azimuth, φ for colatitude, and R for radial. Spherical ...
  11. [11]
    what is the cylindrical coordinate system? - EE-Vibes
    Dec 10, 2023 · **θ (theta)**: Theta still represents the angle measured counterclockwise from the positive x-axis to the projection of the point in the xy- ...
  12. [12]
    [PDF] 12 Cylindrical Coordinates - DigitalCommons@USU
    Warning: Note that ρ > 0 and, without loss of generality, 0 φ < 2π. Cylindrical coordinates are not well behaved on the z-axis. For example, what is φ for a ...
  13. [13]
    Who introduced cylindrical coordinates?
    May 12, 2019 · He introduces what are essentially cylindrical coordinates (r,θ,z) in place of the rectangular coordinates and substitutes into (5)." What Euler ...
  14. [14]
    Cylindrical Coordinates -- from Wolfram MathWorld
    Cylindrical coordinates are a generalization of two-dimensional polar coordinates to three dimensions by superposing a height (z) axis.
  15. [15]
    2.7 Cylindrical and Spherical Coordinates - Calculus Volume 3
    Mar 30, 2016 · In the cylindrical coordinate system, location of a point in space is described using two distances (randz) and an angle measure (θ). In the ...
  16. [16]
  17. [17]
    Cylindrical and spherical coordinates
    Cylindrical coordinates use (r,θ,z) with z as the axis of symmetry. Spherical coordinates use (ρ,θ,ϕ) with the origin at the center of the sphere.
  18. [18]
    [PDF] Cylindrical and Spherical Coordinates
    EX 1 Convert the coordinates as indicated a) (3, π/3, -4) from cylindrical to Cartesian. b) (-2, 2, 3) from Cartesian to cylindrical. Page 5. 5. EX 2 Convert ...
  19. [19]
    9.4 Relations between Cartesian, Cylindrical, and Spherical ...
    The positive z-axes of the cartesian and cylindrical systems coincide with the positive polar axis of the spherical system. The initial rays of the cylindrical ...
  20. [20]
    Calculus III - Spherical Coordinates - Pauls Online Math Notes
    Nov 16, 2022 · Convert the point (√6,π4,√2) ( 6 , π 4 , 2 ) from cylindrical to spherical coordinates. Convert the point (−1,1,−√2) ( − 1 , 1 ...
  21. [21]
    [PDF] Lecture 3: 1.4: Cylindrical and Spherical Coordinates. - UCSD Math
    Cylindrical coordinates (r, θ, z) of a point P(x, y, z) are obtained by using polar coordinates (r, θ) of the projection in the x-y plane and leaving z ...Missing: system | Show results with:system<|control11|><|separator|>
  22. [22]
    [PDF] Coordinate Systems
    dV = dx·dy·dz = r·dr·dθ·dz ds. 2. = dr. 2. + r. 2 dθ 2. + dz. 2. SPHERICAL (page 261) ... Cylindrical coordinates: x1 = r x2 = θ x3 = z. Spherical coordinates ...Missing: line element
  23. [23]
    [PDF] Jacobian - Physics
    Example: Volume element in cylindrical coordinates. • We know that dV = dxdydz in Cartesian coordinates, and also. dV = rdrd cosθdz in cylindrical coordinates ...
  24. [24]
    Triple Integrals in Cylindrical and Spherical Coordinates
    Cylindrical coordinates are obtained from Cartesian coordinates by replacing the x and y coordinates with polar coordinates r and theta and leaving the z ...Missing: multiple representations
  25. [25]
    Calculus III - Triple Integrals in Cylindrical Coordinates
    Nov 16, 2022 · In this section we will look at converting integrals (including dV) in Cartesian coordinates into Cylindrical coordinates.
  26. [26]
    Cylindrical Coordinates - Richard Fitzpatrick
    In the cylindrical coordinate system, $ u_1=r$ , $ u_2=\theta$ , and $ u_3=z$ , where $ r=\sqrt{x^{\,2}+y^{\,Missing: line ds²
  27. [27]
    [PDF] Lecture 5 Vector Operators: Grad, Div and Curl
    The three vector operators are the gradient of a scalar field, the divergence of a vector field, and the curl of a vector field.
  28. [28]
    [PDF] notes.coordinates.pdf - OSU Math
    Cylindrical coordinates (r, θ, z) are related to Cartesian coordinates by r = px² + y², θ = arctan(y/x), z = z. Unit vectors er and eθ are defined.
  29. [29]
    Div, Grad and Curl in Orthogonal Curvilinear Coordinates - Galileo
    The divergence of a vector field →V in curvilinear coordinates is found using Gauss' theorem, that the total vector flux through the six sides of the cube ...
  30. [30]
    [PDF] Physics 504, Lecture 4 Feb. 1, 2010 1 Curvilinear Coordinates
    Feb 1, 2010 · which is a coordinate independent statement, so we have derived the formula for a curl in general orthogonal curvilinear coordinates. 7. When ...
  31. [31]
    6.3. Laplace operator in different coordinates
    The Laplace operator is considered in polar coordinates (r, θ), spherical coordinates (ρ, ϕ, θ), and also in elliptic and parabolic coordinates.
  32. [32]
    Laplace's Equation in Cylindrical Coordinates
    Laplace's equation in cylindrical coordinates is $\displaystyle \nabla^{\,2}\phi = 0, written as (396) in cylindrical coordinates.
  33. [33]
    21. Cylindrical Symmetry: Bessel Functions - Galileo and Einstein
    The Bessel functions are solutions to Laplace's equation in cylindrical coordinates ρ,ϕ,z, that is: ... The usual separation of variables routine leads to ...
  34. [34]
    Separation of variable in polar and cylindrical coordinates
    Consider Laplace equation in the cylinder {r≤a,0≤z≤b} with homogeneous Dirichlet (or Neumann, etc) boundary conditions on the top and bottom leads and non- ...
  35. [35]
    [PDF] Solutions to Laplace's Equation in Cylindrical Coordinates and ...
    Bessel functions oscillate, but are not periodic like harmonic functions, see Figure 1. Thus we expect that the harmonic function solution for. Ψ and the Bessel ...
  36. [36]
    [PDF] Laplace Equation in Cylindrical Coordinates [lam14]
    Bessel equation (left) for x = kr and modified Bessel eq. (right) for x = κr: x2R00(x)+xR0(x)+(x2-ν2)R(x)=0, x2R00(x)+xR0(x)-(x2+ν2)R(x)=0.
  37. [37]
    Laplace's equation in cylindrical coordinates and Bessel's equation (I)
    The method of separation of variables is employed to solve Laplace's equation in cylindrical coordinates. Bessel's equation arises as the r-dependence of the ...<|control11|><|separator|>