A toroid is a doughnut-shaped geometric figure generated by revolving a closed plane curve about an axis in its plane that does not intersect the curve, enclosing a solid volume often resembling a ring or torus when the curve is circular.[1] In physics and engineering, the term commonly refers to a toroidalcoil, consisting of multiple turns of wire wound tightly around a ring-shaped core to form a closed loop, which produces a concentrated magnetic field confined within the coil.[2]Geometrically, the toroid generalizes the torus—the specific case where a circle is rotated about an axis in its plane at a fixed distance from the center—yielding a surface of revolution with applications in topology, where it serves as a fundamental non-spherical shape.[1] The volume and surface area of a toroidal solid can be calculated using integrals based on the generating curve, with the standard torus having volume V = 2\pi^2 R r^2 and surface area A = 4\pi^2 R r, where R is the distance from the center of the tube to the center of the torus and r is the tube radius.[1]In electromagnetism, a toroid functions as a bent solenoid, offering cylindrical symmetry that results in a magnetic field \mathbf{B} inside the coil approximately uniform and equal to B = \frac{\mu_0 N I}{2\pi R} (where \mu_0 is the permeability of free space, N is the number of turns, I is the current, and R is the mean radius), while the field outside is ideally zero due to the closed path enclosing no net current for external Amperian loops.[2] This containment of the field minimizes leakage and interference, making toroids superior to straight solenoids for applications requiring efficient energy storage and transfer.[3]Toroidal coils are widely employed in electrical engineering as inductors and transformers due to their compact design, low electromagnetic interference, and high efficiency.[4][3] Key uses include power supplies, audio amplifiers, and medical imaging devices like MRI machines, where they enable precise current sensing and voltage stepping without external field disruptions.[4] In advanced physics, toroidal geometries model plasma confinement in fusion reactors, such as tokamaks, leveraging their ability to sustain stable, looping magnetic fields.[5]
Definition and Basic Concepts
Mathematical Definition
In mathematics, a toroid is defined as a surface of revolution obtained by rotating a closed plane curve about an axis lying in its plane that does not intersect the curve, thereby generating a surface featuring a central hole.[1] This construction distinguishes the toroid from simpler surfaces like spheres or cylinders, as the non-intersecting axis ensures the resulting shape encloses an empty interior region. The generating curve is typically a circle or a polygon, with the circle yielding the most common form known as the torus.[1]The key parameters characterizing a standard toroid are the major radius R, defined as the distance from the axis of rotation to the center of the generating curve, and the minor radius r, the radius of the generating curve itself. For a non-self-intersecting ring-shaped toroid, the condition R > r must hold, preventing the surface from crossing itself.[6] Visually, this produces a doughnut-like shape with a genus of 1, meaning it possesses exactly one hole, a topological property that underscores its fundamental role in surface classification.[6]The term "toroid" was coined in the late 19th century, with its first recorded use in 1886, to generalize surfaces beyond the specific circular case of the torus.[7] For the basic toroid surface generated by a circle, the implicit equation in three-dimensional Cartesian coordinates is given by\left( \sqrt{x^2 + y^2} - R \right)^2 + z^2 = r^2,where the axis of rotation aligns with the z-axis.[6] This equation encapsulates the surface's geometry, with points satisfying it lying on the toroid.
Relation to the Torus
The torus represents a specific instance of a toroid, generated by revolving a circle of constantradius (the minorradius) around an axis in its plane that lies outside the circle itself, resulting in a surface of revolution with uniform tube thickness.[1][6] This configuration distinguishes the torus from more general toroids, where the generating curve may deviate from a circle, leading to variable cross-sections.Topologically, the torus and the standard toroid share identical properties, including an Euler characteristic of 0 and a genus of 1, which characterize them as orientable surfaces with a single hole.[8] These invariants ensure that the standard torus can be embedded in three-dimensional Euclidean space without self-intersection, provided the major radius exceeds the minor radius to avoid overlap.[6]In mathematical usage, the term "torus" is typically reserved for this ideal circular surface, particularly in pure mathematics and topology, where it serves as the fundamental example of a genus-1 surface.[6] In contrast, "toroid" encompasses a broader class, including non-circular surfaces of revolution, solid volumes with ring-like topology (such as the solid torus), and even polyhedral approximations that are topologically equivalent to a torus.[1][9] This distinction arises from the etymology: "torus" derives from the Latin torus, meaning "swelling," "bulge," or "cushion," originally referring to architectural moldings and adopted in mathematics around the 16th century; "toroid," coined later in 1886 from New Latin roots, emphasizes the ring-like (-oid) form more generally.[10][7]A canonical example of the torus appears in toroidal coordinates (\eta, \tau, \phi), where surfaces of constant \tau define toroidal surfaces, with the standard torus given by \tau = constant, parameterized as:\begin{align*}
x &= \frac{a \sinh \tau \cos \phi}{\cosh \tau - \cos \eta}, \\
y &= \frac{a \sinh \tau \sin \phi}{\cosh \tau - \cos \eta}, \\
z &= \frac{a \sin \eta}{\cosh \tau - \cos \eta},
\end{align*}where a > 0 is a scale factor, \eta \in [0, 2\pi), \tau \in [0, \infty), and \phi \in [0, 2\pi).[11]
Geometric Properties
Parametric Equations
The parametric equations for the surface of a standard toroid, generated by revolving a circle of radius r (the minor radius) centered at (R, 0, 0) in the xz-plane around the z-axis, where R is the major radius, are given by\begin{align*}
x &= (R + r \cos \theta) \cos \phi, \\
y &= (R + r \cos \theta) \sin \phi, \\
z &= r \sin \theta,
\end{align*}with angular parameters \theta, \phi \in [0, 2\pi).[6] These equations parameterize points on the toroidal surface, enabling computational generation and visualization in three-dimensional space.[6]The derivation begins with the parametric form of the generating circle in the xz-plane: x = R + r \cos \theta, z = r \sin \theta, y = 0. Revolving this circle around the z-axis introduces the azimuthal angle \phi, transforming the x- and y-coordinates via rotation: x' = x \cos \phi, y' = x \sin \phi, while z remains unchanged, yielding the full parametric equations.[12]An equivalent implicit equation for the circular toroid, eliminating the parameters, is\left( \sqrt{x^2 + y^2} - R \right)^2 + z^2 = r^2.This form describes the set of points at a fixed distance r from a circle of radius R in the xy-plane.[6]For non-circular generating curves, the parametric equations adapt by replacing the circular terms with the appropriate parametrization of the cross-section. For an elliptical cross-section with semi-major axis a along x and semi-minor axis b along z, the equations become x = (R + a \cos \theta) \cos \phi, y = (R + a \cos \theta) \sin \phi, z = b \sin \theta.[13] Similarly, for polygonal paths, the generating curve can be parameterized piecewise using linear segments between vertices of the polygon, then revolved around the axis to form the toroid surface.[14]The shape of the toroid varies with the ratio of R to r. When R > r, the surface forms a non-self-intersecting ring toroid. As R approaches r, the inner equator pinches to a point, creating a horn toroid at R = r. For R < r, self-intersections occur along the inner surface, resulting in a spindle toroid. For example, with R = 2 and r = 1, the surface remains a smooth ring without overlap; but with R = 1 and r = 2, the parameters produce a self-intersecting spindle shape.[6]
Surface Area and Volume Calculations
The surface area and volume of a toroid are computed using its parametric equations, which provide the foundation for integration over the surface or solid. For the standard ring toroid, generated by revolving a circle of radius r (tube radius) around an axis at distance R (major radius) from its center, with R > r, explicit closed-form formulas exist. These derive from either Pappus's centroid theorem or direct surface and volume integrals.Pappus's centroid theorem, attributed to the Greek mathematician Pappus of Alexandria (c. 290–350 AD), simplifies the calculations for surfaces and solids of revolution. For the surface area S, the theorem states that S equals the arc length of the generating curve times the distance traveled by its centroid. The generating circle has arc length $2\pi r, and its centroid travels a circular path of radius R with length $2\pi R, yielding S = (2\pi r)(2\pi R) = 4\pi^2 R r. Similarly, for the volume V of the solid toroid, the theorem uses the area of the generating disk \pi r^2 and the same centroid path length, giving V = (\pi r^2)(2\pi R) = 2\pi^2 R r^2.Direct derivations via calculus confirm these results and offer insight into the geometry. The parametric equations for the torus surface are \mathbf{r}(\theta, \phi) = ((R + r \cos \theta) \cos \phi, (R + r \cos \theta) \sin \phi, r \sin \theta), where $0 \leq \theta, \phi < 2\pi. The surface area element is dS = \| \mathbf{r}_\theta \times \mathbf{r}_\phi \| \, d\theta \, d\phi = r (R + r \cos \theta) \, d\theta \, d\phi. Integrating over the parameter domain gives S = \int_0^{2\pi} \int_0^{2\pi} r (R + r \cos \theta) \, d\theta \, d\phi = 4\pi^2 R r. For the volume, methods like the washer or shell integration, or the divergence theorem applied to the solid, yield V = 2\pi^2 R r^2; Pappus provides an equivalent non-integral approach emphasizing the centroid's role.These formulas assume a circular cross-section and R > r to ensure a non-self-intersecting ring shape without a hole closure. For non-circular toroids, where the generating curve is an arbitrary closed shape, closed forms are unavailable, and the surface area requires numerical evaluation of the general surface integral S = \iint_D \| \mathbf{r}_u \times \mathbf{r}_v \| \, du \, dv over the parameter domain D, often using quadrature methods or discretization. Volume computations similarly rely on numerical integration or Monte Carlo techniques for the enclosed region. When R < r, as in spindle toroids, the surface self-intersects, invalidating the standard formulas for the enclosed volume; modifications, such as subtracting overlapping regions or using inclusion-exclusion, are needed to compute the actual solid volume accurately.
Types and Variations
Ring Torus
The ring torus, also known as the standard or common torus, is generated by revolving a circle of radius r about an axis coplanar with the circle but displaced by a distance R from its center, where R > r to ensure the surface does not self-intersect.[6] This condition prevents the generating circle from crossing the axis of rotation, producing a smooth, doughnut-like ring without overlaps or singularities.[15] The resulting surface is a surface of revolution characterized by its rotational symmetry around the central axis.Geometrically, the ring torus features a tube whose centerline maintains a constant distance R from the axis of symmetry, forming a circular path of major radius R. The cross-section perpendicular to this centerline is uniformly circular with radius r, ensuring consistent thickness throughout the structure. This uniformity contributes to the ring torus's topological genus of 1, distinguishing it as a compact orientable surface with a single hole. Visually, the absence of intersections or cusps renders it stable and aesthetically simple, ideal for physical prototypes and simulations where smooth curvature is essential.[6][16]A key special property of the ring torus is its central role in standard toroidal coordinates, a curvilinear system where surfaces of constant coordinate values describe toroidal shapes aligned with the ring geometry. These coordinates facilitate analysis in fields requiring axial symmetry, such as potential theory and fluid dynamics around toroidal objects. Regarding mappings, certain abstract representations of the torus admit isometric immersions into higher-dimensional spheres under specific metric conditions, though the embedded ring torus in Euclidean space exhibits varying Gaussian curvature.[11][17]For quantitative aspects, substituting the ring torus parameters into general surface area and volume formulas—typically derived via integration or Pappus's centroid theorem—yields precise closed-form results: the surface area is $4\pi^2 R r, and the enclosed volume of the solid formis $2\pi^2 R r^2. These expressions highlight the scalability of the shape, with area growing linearly in both radii and volume quadratically in the minor radius.[6]
Spindle and Horn Tori
Spindle tori and horn tori represent degenerate cases of the standard torus generated by revolving a circle of radius r around an axis at distance R from the circle's center, where R \leq r. In these configurations, the generating circle intersects or touches the axis of revolution, leading to self-intersections or singularities that distinguish them from the non-degenerate ring torus.[18]The spindle torus arises when R < r, causing the revolved surface to self-intersect and form a characteristic "apple" shape. The self-intersection occurs at two singular points located on the axis of rotation, symmetric about the equatorial plane, where the surface crosses itself without a well-defined tangent plane. These points correspond to the locations where the generating circle intersects the axis before revolution.[19]The horn torus forms when R = r, with the generating circle tangent to the axis at a single point. This results in a horn-like cusp singularity at the origin, where the inner equator degenerates to a point and the surface develops a conical pinch without a tangent plane. The cusp manifests as a sharp spike directed toward the axis.[19]The parametric equations for both spindle and horn tori are identical to those of the ring torus:\begin{align*}
x &= (R + r \cos \theta) \cos \phi, \\
y &= (R + r \cos \theta) \sin \phi, \\
z &= r \sin \theta,
\end{align*}where \theta, \phi \in [0, 2\pi). For R \leq r, the radial term R + r \cos \theta becomes non-positive for certain \theta, specifically zero when \cos \theta = -R/r. This yields self-intersection points at (x, y, z) = (0, 0, \pm r \sqrt{1 - (R/r)^2}) for the spindle torus. At the bifurcation point R = r, these two points coincide at the origin, transitioning from double singularities to a single cusp. This parametric adjustment highlights a geometric bifurcation as R decreases through r, closing the central hole and introducing singularities.[18][19]Geometrically, the axis intersection alters the embedding while preserving the topological genus of 1, equivalent to a smooth torus, though the resulting surfaces are immersed with singularities that prevent them from being smooth manifolds. In the spindle torus, the self-intersections divide the surface into inner ("lemon") and outer ("apple") regions connected at the singular points, confining geodesics to disjoint components. The horn torus features a single connected surface with the cusp acting as an infinite potential barrier for geodesics, eliminating unbound paths. These properties make spindle and horn tori valuable in analyzing singularity formation and geodesic behavior on surfaces of revolution.[19]
Square and Other Non-Circular Toroids
A square toroid is generated by revolving a square closed curve around an axis lying in the plane of the square but external to it, typically parallel to two opposite sides, yielding a surface composed of four distinct portions: two cylindrical surfaces from the sides parallel to the axis and two annular disk-like surfaces from the perpendicular sides. This configuration produces a frame-like structure with sharp edges at the corners of the original square, distinguishing it from the smooth circular torus.[20]The parametric representation for a square toroid requires a piecewise definition corresponding to the four linear segments of the square generator. For a square of side length s centered at distance R from the revolution axis (with R > s/2), the parameter \theta (revolution angle, $0 \leq \theta < 2\pi) is used alongside a segment parameter \phi (0 to 1 for each side). For the bottom side (z = -s/2, x from R - s/2 to R + s/2), the equations are x = (R + (s \phi - s/2)) \cos \theta, y = (R + (s \phi - s/2)) \sin \theta, z = -s/2; similar linear adjustments apply to the other sides by shifting in z or x. This piecewise linear form in the generator parameter introduces discontinuities in the first derivative at corners, complicating differentiability compared to circular cases.[20]Properties of the square toroid exhibit increased complexity in curvature distribution, with zero Gaussian curvature along the cylindrical portions and concentrated strains at the revolved corners, often requiring approximations like filleting in practical models. Volume calculations leverage Pappus's centroid theorem, yielding V = 2\pi R s^2 for a solid square toroid, where R is the centroid distance (mean radius) and s^2 the cross-sectional area, simplifying integration over the non-smooth boundary. Surface area approximations sum contributions from each revolved segment: $2\pi (R - s/2) s + 2\pi (R + s/2) s + 2 \times 2\pi R s for the two cylinders and two annuli, respectively, though exact values demand careful handling of edge effects.Other non-circular toroids include elliptical variants, formed by revolving an ellipse around an external axis coplanar with it. An elliptical toroid has parametric equations x = (c + a \cos v) \cos u, y = (c + a \cos v) \sin u, z = b \sin v, where u, v \in [0, 2\pi), c > a is the distance from the axis to the ellipse center, a and b are the semi-major and semi-minor axes of the elliptical cross-section (with eccentricity e = \sqrt{1 - (b/a)^2} for a > b). Increasing eccentricity stretches the cross-section, altering the surface's aspect ratio and enhancing asymmetry in curvature, which affects metrics like mean curvature more variably than in circular toroids. Volume follows Pappus's theorem as V = 2\pi c \cdot \pi a b, emphasizing the role of the elliptical area \pi a b.[13]Historically, square and other non-circular toroids served as simplified models in early engineering for toroidal coils and structural frames, facilitating manual calculations of magnetic fields or stresses before computational geometry tools enabled precise circular designs; for instance, square cross-sections approximated uniform flux in pre-digital inductor prototypes.[21]
In electromagnetism, a toroid serves as the core for a toroidal solenoid, formed by winding insulated wire uniformly around a ring-shaped structure to generate a magnetic field confined primarily within the interior. This configuration exploits the closed-loop geometry of the toroid to produce a nearly uniform magnetic field inside the core while minimizing the field outside, approaching zero due to the symmetric cancellation of flux lines.[22][23]The concept of the toroidal solenoid traces its origins to the early 19th century, when Michael Faraday developed the first practical electromagnetic induction device in 1831, known as the "induction ring," which consisted of an iron ring wound with two separate coils to demonstrate mutual induction. This invention laid the groundwork for efficient inductive devices by showing how a changing current in one coil could induce voltage in another without direct electrical connection.[24][25]For a tightly wound toroidal solenoid carrying current I with N total turns, the magnetic field strength B at a radial distance r from the center (where r is between the inner and outer radii of the toroid) is given by Ampère's law asB = \frac{\mu_0 N I}{2\pi r},where \mu_0 is the permeability of free space; this field varies inversely with r but remains azimuthal and contained within the core for an ideal case.[22][23]Toroidal solenoids find primary applications in transformers and inductors, where the primary winding carries the input current and a secondary winding captures the induced voltage, enabling efficient power transfer across voltage levels. Compared to linear solenoids, toroidal designs offer significant advantages, including reduced magnetic leakage flux—often less than 1% outside the core—due to the closed magnetic path, which minimizes electromagnetic interference and improves energy efficiency in compact devices.[26][27]To enhance performance, toroidal cores are typically constructed from ferromagnetic materials such as laminated iron, ferrite, or powdered iron, which increase the relative permeability \mu_r (often 100–10,000 times that of air) and thereby amplify the magnetic field and inductance without saturation at typical operating currents. The self-inductance L of such a toroid, assuming a thin ring with mean radius R and cross-sectional area A, is approximated byL = \frac{\mu_0 N^2 A}{2\pi R},where the formula scales with N^2 to reflect the mutual reinforcement of flux from multiple turns; for ferromagnetic cores, \mu_0 is replaced by \mu = \mu_r \mu_0 to account for the material's enhancement.[28][29]
In Plasma Physics and Fusion
In plasma physics and fusion research, toroids play a central role in magnetic confinement devices designed to sustain high-temperature plasmas for controlled nuclear fusion. The toroidal geometry, resembling a doughnut shape, allows for the creation of nested magnetic flux surfaces that prevent plasma particles from escaping due to the Lorentz force acting on charged ions and electrons. This configuration is essential for achieving the extreme conditions required for fusion reactions, where plasma temperatures exceed 100 million degrees Celsius and densities must be maintained long enough for deuterium-tritium nuclei to fuse. Devices like tokamaks and stellarators exploit this shape to generate helical magnetic field lines that wrap around the plasma, providing stability against both radial and axial losses.[30][31]Tokamaks confine plasma using a combination of toroidal and poloidal magnetic fields: the toroidal field B_{\tor} is produced by external coils encircling the torus, while the poloidal field B_{\pol} arises from an induced plasma current flowing toroidally. This interplay creates helical field lines, characterized by the safety factor q = \frac{r B_{\tor}}{R B_{\pol}}, where r is the minor radius and R is the major radius of the toroid; q represents the number of toroidal turns a field line makes per poloidal turn and must typically exceed 2 at the plasma edge to avoid disruptive instabilities. Stellarators, in contrast, achieve similar helical confinement without relying on plasma currents, using twisted external coils to produce both field components directly, which enhances steady-state operation but increases design complexity. The toroidal curvature introduces geometric effects into the magnetohydrodynamic (MHD) equations, such as the Shafranov shift, which displaces the plasma equilibrium away from the geometric center and influences pressure gradients.[32][33]Key challenges in toroidal confinement include MHD instabilities driven by the non-uniformity of the toroid. Kink instabilities occur when the safety factor [q](/page/Q) falls below critical values, causing helical deformations of the plasma column that can lead to sudden reconfiguration or disruption. Ballooning instabilities, exacerbated by the toroid's varying curvature (stronger on the inner side), arise from adverse pressure gradients in regions of weak magnetic shear, potentially ejecting plasma blobs outward. These modes are analyzed through the ideal MHD energyprinciple, where toroidal effects modify the stability criteria compared to cylindrical approximations. Historical milestones trace back to the Soviet Union in the 1950s, where Igor Tamm and Andrei Sakharov proposed the tokamak concept, leading to the first operational device, T-1, in 1958 at the Kurchatov Institute, which demonstrated plasma confinement times of milliseconds.[34][35][36]The ITER project exemplifies modern toroidal fusion efforts, featuring a large toroidal vacuum chamber with a major radius of 6.2 meters to house a plasma volume of 830 cubic meters, aiming to produce 500 MW of fusion power from 50 MW input, with first plasma now scheduled for 2034–2035 as of 2025. This international endeavor integrates advanced superconducting magnets to sustain fields up to 13 tesla, addressing scalability issues in prior devices; recent milestones include the completion of the central solenoid in 2025.[30][37][38] Related compact toroids, such as spheromaks and field-reversed configurations (FRCs), lack a central column; FRCs are prolate configurations that reverse the field at the edge while maintaining poloidal dominance inside, enabling higher beta values (plasma pressure over magnetic pressure) for efficient confinement in smaller volumes. Spheromaks are formed by magnetized plasma guns and exhibit force-free equilibria with both toroidal and poloidal fields, though they face tilt instabilities absent in axisymmetric tokamaks.[39]