Everipedia Logo
Everipedia is now IQ.wiki - Join the IQ Brainlist and our Discord for early access to editing on the new platform and to participate in the beta testing.
Angular momentum

Angular momentum

In physics, angular momentum (rarely, moment of momentum or rotational momentum) is the rotational equivalent of linear momentum. It is an important quantity in physics because it is a conserved quantity—the total angular momentum of a closed system remains constant.

In three dimensions, the angular momentum for a point particle is a pseudovector r × p, the cross product of the particle's position vector r (relative to some origin) and its momentum vector; the latter is p = mv in Newtonian mechanics. This definition can be applied to each point in continua like solids or fluids, or physical fields. Unlike momentum, angular momentum does depend on where the origin is chosen, since the particle's position is measured from it.

Just like for angular velocity, there are two special types of angular momentum: the spin angular momentum and the orbital angular momentum. The spin angular momentum of an object is defined as the angular momentum about its centre of mass coordinate. The orbital angular momentum of an object about a chosen origin is defined as the angular momentum of the centre of mass about the origin. The total angular momentum of an object is the sum of the spin and orbital angular momenta. The orbital angular momentum vector of a particle is always parallel and directly proportional to the orbital angular velocity vector ω of the particle, where the constant of proportionality depends on both the mass of the particle and its distance from origin. However, the spin angular momentum of the object is proportional but not always parallel to the spin angular velocity Ω, making the constant of proportionality a second-rank tensor rather than a scalar.

Angular momentum is additive; the total angular momentum of any composite system is the (pseudo) vector sum of the angular momenta of its constituent parts. For a continuous rigid body, the total angular momentum is the volume integral of angular momentum density (i.e. angular momentum per unit volume in the limit as volume shrinks to zero) over the entire body.

Torque can be defined as the rate of change of angular momentum, analogous to force. The net external torque on any system is always equal to the total torque on the system; in other words, the sum of all internal torques of any system is always 0 (this is the rotational analogue of Newton's Third Law). Therefore, for a closed system (where there is no net external torque), the total torque on the system must be 0, which means that the total angular momentum of the system is constant. The conservation of angular momentum helps explain many observed phenomena, for example the increase in rotational speed of a spinning figure skater as the skater's arms are contracted, the high rotational rates of neutron stars, the Coriolis effect, and the precession of gyroscopes. In general, conservation does limit the possible motion of a system, but does not uniquely determine what the exact motion is.

In quantum mechanics, angular momentum (like other quantities) is expressed as an operator, and its one-dimensional projections have quantized eigenvalues. Angular momentum is subject to the Heisenberg uncertainty principle, implying that at any time, only one projection (also called "component") can be measured with definite precision; the other two then remain uncertain. Because of this, it turns out that the notion of a quantum particle literally "spinning" about an axis does not exist. Nevertheless, elementary particles still possess a spin angular momentum, but this angular momentum does not correspond to spinning motion in the ordinary sense.[1]

Angular momentum
Common symbols
L
In SI base unitskg m2s−1
Conserved?yes
Derivations from
other quantities
L = Iω = r × p
DimensionM L2T−1

In classical mechanics

Definition

Orbital angular momentum in two dimensions

Velocity of the particle m with respect to the origin O can be resolved into components parallel to (v∥) and perpendicular to (v⊥) the radius vector r. The angular momentum of m is proportional to the perpendicular component v⊥ of the velocity, or equivalently, to the perpendicular distance r⊥ from the origin.

Velocity of the particle m with respect to the origin O can be resolved into components parallel to (v∥) and perpendicular to (v⊥) the radius vector r. The angular momentum of m is proportional to the perpendicular component v⊥ of the velocity, or equivalently, to the perpendicular distance r⊥ from the origin.

Angular momentum is avectorquantity (more precisely, apseudovector) that represents the product of a body'srotational inertiaandrotational velocity(in radians/sec) about a particular axis. However, if the particle's trajectory lies in a singleplane, it is sufficient to discard the vector nature of angular momentum, and treat it as ascalar(more precisely, apseudoscalar).[2] Angular momentum can be considered a rotational analog oflinear momentum. Thus, wherelinear momentumis proportional tomassandlinear speed,
angular momentumis proportional tomoment of inertiaandangular speedmeasured in radians per second.[3]
Unlike mass, which depends only on amount of matter, moment of inertia is also dependent on the position of the axis of rotation and the shape of the matter. Unlike linear speed, which does not depend upon the choice of origin, angular velocity is always measured with respect to a fixed origin. Therefore, strictly speaking,should be referred to as the angular momentum relative to that center.[4]
Becausefor a single particle andfor circular motion, angular momentum can be expanded,and reduced to,
the product of theradiusof rotationand thelinear momentumof the particle, wherein this case is the equivalentlinear (tangential) speedat the radius ().

This simple analysis can also apply to non-circular motion if only the component of the motion which is perpendicular to the radius vector is considered. In that case,

whereis the perpendicular component of the motion. Expanding,rearranging,and reducing, angular momentum can also be expressed,
whereis the length of themoment arm, a line dropped perpendicularly from the origin onto the path of the particle. It is this definition,(length of moment arm)×(linear momentum)to which the term moment of momentum refers.[5]

Scalar—angular momentum from Lagrangian mechanics

Another approach is to define angular momentum as the conjugate momentum (also called canonical momentum) of the angular coordinateexpressed in theLagrangianof the mechanical system. Consider a mechanical system with a massconstrained to move in a circle of radiusin the absence of any external force field. The kinetic energy of the system is

And the potential energy is

Then the Lagrangian is

The generalized momentum "canonically conjugate to" the coordinateis defined by

Orbital angular momentum in three dimensions

Relationship between force (F), torque (τ), momentum (p), and angular momentum (L) vectors in a rotating system. r is the position vector.

Relationship between force (F), torque (τ), momentum (p), and angular momentum (L) vectors in a rotating system. r is the position vector.

To completely define orbital angular momentum in three dimensions, it is required to know the rate at which the position vector sweeps out angle, the direction perpendicular to the instantaneous plane of angular displacement, and the mass involved, as well as how this mass is distributed in space[6]. By retaining this vector nature of angular momentum, the general nature of the equations is also retained, and can describe any sort of three-dimensional motion about the center of rotation – circular, linear, or otherwise. In vector notation, the orbital angular momentum of a point particle in motion about the origin can be expressed as:

where

is themoment of inertiafor apoint mass,
is the orbitalangular velocityin radians/sec (units 1/sec) of the particle about the origin,
is theposition vectorof the particle relative to the origin,,
is thelinear velocityof the particle relative to the origin, and
is themassof the particle.

This can be expanded, reduced, and by the rules of vector algebra, rearranged:

which is thecross productof the position vectorand the linear momentumof the particle. By the definition of the cross product, thevector isperpendicularto bothand. It is directed perpendicular to the plane of angular displacement, as indicated by theright-hand rule– so that the angular velocity is seen ascounter-clockwisefrom the head of the vector. Conversely, thevector defines theplanein whichandlie.
By defining aunit vectorperpendicular to the plane of angular displacement, ascalar angular speedresults, where
andwhereis the perpendicular component of the motion, as above.

The two-dimensional scalar equations of the previous section can thus be given direction:

andfor circular motion, where all of the motion is perpendicular to the radius.

Discussion

Moment of inertia (shown here), and therefore angular momentum, is different for every possible configuration of mass and axis of rotation.

Moment of inertia (shown here), and therefore angular momentum, is different for every possible configuration of mass and axis of rotation.

Angular momentum can be described as the rotational analog of linear momentum. Like linear momentum it involves elements of mass and displacement. Unlike linear momentum it also involves elements of position and shape.

Many problems in physics involve matter in motion about some certain point in space, be it in actual rotation about it, or simply moving past it, where it is desired to know what effect the moving matter has on the point—can it exert energy upon it or perform work about it? Energy, the ability to do work, can be stored in matter by setting it in motion—a combination of its inertia and its displacement. Inertia is measured by its mass, and displacement by its velocity. Their product,

is the matter's momentum.[7] Referring this momentum to a central point introduces a complication: the momentum is not applied to the point directly. For instance, a particle of matter at the outer edge of a wheel is, in effect, at the end of a lever of the same length as the wheel's radius, its momentum turning the lever about the center point. This imaginary lever is known as the moment arm. It has the effect of multiplying the momentum's effort in proportion to its length, an effect known as a moment. Hence, the particle's momentum referred to a particular point,

is the angular momentum, sometimes called, as here, the moment of momentum of the particle versus that particular center point. The equationcombines a moment (a massturning moment arm) with a linear (straight-line equivalent) speed. Linear speed referred to the central point is simply the product of the distanceand the angular speedversus the point:another moment. Hence, angular momentum contains a double moment:Simplifying slightly,the quantityis the particle'smoment of inertia, sometimes called the second moment of mass. It is a measure of rotational inertia.[8]

Because moment of inertia is a crucial part of the spin angular momentum, the latter necessarily includes all of the complications of the former, which is calculated by multiplying elementary bits of the mass by the squares of their distances from the center of rotation.[9] Therefore, the total moment of inertia, and the angular momentum, is a complex function of the configuration of the matter about the center of rotation and the orientation of the rotation for the various bits.

For a rigid body, for instance a wheel or an asteroid, the orientation of rotation is simply the position of the rotation axis versus the matter of the body. It may or may not pass through the center of mass, or it may lie completely outside of the body. For the same body, angular momentum may take a different value for every possible axis about which rotation may take place.[10] It reaches a minimum when the axis passes through the center of mass.[11]

For a collection of objects revolving about a center, for instance all of the bodies of the Solar System, the orientations may be somewhat organized, as is the Solar System, with most of the bodies' axes lying close to the system's axis. Their orientations may also be completely random.

In brief, the more mass and the farther it is from the center of rotation (the longer the moment arm), the greater the moment of inertia, and therefore the greater the angular momentum for a given angular velocity. In many cases the moment of inertia, and hence the angular momentum, can be simplified by,[12]

whereis theradius of gyration, the distance from the axis at which the entire massmay be considered as concentrated.
Similarly, for apoint massthemoment of inertiais defined as,
whereis theradiusof the point mass from the center of rotation,
and for any collection of particlesas the sum,

Angular momentum's dependence on position and shape is reflected in its units versus linear momentum: kg⋅m2/s, N⋅m⋅s, or J⋅s for angular momentum versus kg⋅m/s or N⋅s for linear momentum. When calculating angular momentum as the product of the moment of inertia times the angular velocity, the angular velocity must be expressed in radians per second, where the radian assumes the dimensionless value of unity. (When performing dimensional analysis, it may be productive to use orientational analysis which treats radians as a base unit, but this is outside the scope of the International system of units). Angular momentum's units can be interpreted as torque⋅time or as energy⋅time per angle. An object with angular momentum of L N⋅m⋅s can be reduced to zero rotation (all of the rotational energy can be transferred out of it) by an angular impulse of L N⋅m⋅s[13] or equivalently, by torque or work of L N⋅m for one second, or energy of L J for one second.[14]

The plane perpendicular to the axis of angular momentum and passing through the center of mass[15] is sometimes called the invariable plane, because the direction of the axis remains fixed if only the interactions of the bodies within the system, free from outside influences, are considered.[16] One such plane is the invariable plane of the Solar System.

Angular momentum and torque

Newton's second law of motion can be expressed mathematically,

or force = mass × acceleration. The rotational equivalent for point particles may be derived as follows:

which means that the torque (i.e. the time derivative of the angular momentum) is

Because the moment of inertia is, it follows that, andwhich, reduces to

This is the rotational analog of Newton's Second Law. Note that the torque is not necessarily proportional or parallel to the angular acceleration (as one might expect). The reason for this is that the moment of inertia of a particle can change with time, something that cannot occur for ordinary mass.

Conservation of angular momentum

A figure skater conserves angular momentum – her rotational speed increases as her moment of inertia decreases by drawing in her arms and legs.

A figure skater conserves angular momentum – her rotational speed increases as her moment of inertia decreases by drawing in her arms and legs.

The torque caused by the two opposing forces Fg and −Fg causes a change in the angular momentum L in the direction of that torque (since torque is the time derivative of angular momentum). This causes the top to precess.

The torque caused by the two opposing forces Fg and −Fg causes a change in the angular momentum L in the direction of that torque (since torque is the time derivative of angular momentum). This causes the top to precess.

A rotational analog of Newton's third law of motion might be written, "In a closed system, no torque can be exerted on any matter without the exertion on some other matter of an equal and opposite torque."[17] Hence, angular momentum can be exchanged between objects in a closed system, but total angular momentum before and after an exchange remains constant (is conserved).[18]

Seen another way, a rotational analogue of Newton's first law of motion might be written, "A rigid body continues in a state of uniform rotation unless acted by an external influence."[17] Thus with no external influence to act upon it, the original angular momentum of the system remains constant.[19]

The conservation of angular momentum is used in analyzingcentral force motion. If the net force on some body is directed always toward some point, the center, then there is no torque on the body with respect to the center, as all of the force is directed along theradius vector, and none isperpendicularto the radius. Mathematically, torquebecause in this caseandare parallel vectors. Therefore, the angular momentum of the body about the center is constant. This is the case withgravitational attractionin theorbitsofplanetsandsatellites, where the gravitational force is always directed toward the primary body and orbiting bodies conserve angular momentum by exchanging distance and velocity as they move about the primary. Central force motion is also used in the analysis of theBohr modelof theatom.

For a planet, angular momentum is distributed between the spin of the planet and its revolution in its orbit, and these are often exchanged by various mechanisms. The conservation of angular momentum in the Earth–Moon system results in the transfer of angular momentum from Earth to Moon, due to tidal torque the Moon exerts on the Earth. This in turn results in the slowing down of the rotation rate of Earth, at about 65.7 nanoseconds per day,[20] and in gradual increase of the radius of Moon's orbit, at about 3.82 centimeters per year.[21]

The conservation of angular momentum explains the angular acceleration of an ice skater as she brings her arms and legs close to the vertical axis of rotation. By bringing part of the mass of her body closer to the axis, she decreases her body's moment of inertia. Because angular momentum is the product of moment of inertia and angular velocity, if the angular momentum remains constant (is conserved), then the angular velocity (rotational speed) of the skater must increase.

The same phenomenon results in extremely fast spin of compact stars (like white dwarfs, neutron stars and black holes) when they are formed out of much larger and slower rotating stars. Decrease in the size of an object n times results in increase of its angular velocity by the factor of n2.

Conservation is not always a full explanation for the dynamics of a system but is a key constraint. For example, a spinning top is subject to gravitational torque making it lean over and change the angular momentum about the nutation axis, but neglecting friction at the point of spinning contact, it has a conserved angular momentum about its spinning axis, and another about its precession axis. Also, in any planetary system, the planets, star(s), comets, and asteroids can all move in numerous complicated ways, but only so that the angular momentum of the system is conserved.

Noether's theorem states that every conservation law is associated with a symmetry (invariant) of the underlying physics. The symmetry associated with conservation of angular momentum is rotational invariance. The fact that the physics of a system is unchanged if it is rotated by any angle about an axis implies that angular momentum is conserved.[22]

Angular momentum in orbital mechanics

In astrodynamics and celestial mechanics, a massless (or per unit mass) angular momentum is defined[23]

called *specific angular momentum*. Note thatMassis often unimportant in orbital mechanics calculations, because motion is defined bygravity. The primary body of the system is often so much larger than any bodies in motion about it that the smaller bodies have a negligible gravitational effect on it; it is, in effect, stationary. All bodies are apparently attracted by its gravity in the same way, regardless of mass, and therefore all move approximately the same way under the same conditions.

Solid bodies

For a continuous mass distribution with density function ρ(r), a differential volume element dV with position vector r within the mass has a mass element dm = ρ(r)dV. Therefore, the infinitesimal angular momentum of this element is:

and integrating this differential over the volume of the entire mass gives its total angular momentum:

In the derivation which follows, integrals similar to this can replace the sums for the case of continuous mass.

Collection of particles

Center of mass

The angular momentum of the particles i is the sum of the cross products R × MV + Σri × mivi.

The angular momentum of the particles i is the sum of the cross products R × MV + Σri × mivi.

For a collection of particles in motion about an arbitrary origin, it is informative to develop the equation of angular momentum by resolving their motion into components about their own center of mass and about the origin. Given,

is the mass of particle,is the position vector of particlevs the origin,is the velocity of particlevs the origin,is the position vector of the center of mass vs the origin,is the velocity of the center of mass vs the origin,is the position vector of particlevs the center of mass,is the velocity of particlevs the center of mass,

The total mass of the particles is simply their sum,

The position vector of the center of mass is defined by,[24]

By inspection,

and

The total angular momentum of the collection of particles is the sum of the angular momentum of each particle,

Expanding,
Expanding,

It can be shown that (see sidebar),

Prove thatwhich, by the definition of the center of mass, isand similarly for
and

therefore the second and third terms vanish,

The first term can be rearranged,

and total angular momentum for the collection of particles is finally,[25]

The first term is the angular momentum of the center of mass relative to the origin. Similar to Single particle, below, it is the angular momentum of one particle of mass M at the center of mass moving with velocity V. The second term is the angular momentum of the particles moving relative to the center of mass, similar to Fixed center of mass, below. The result is general—the motion of the particles is not restricted to rotation or revolution about the origin or center of mass. The particles need not be individual masses, but can be elements of a continuous distribution, such as a solid body.

Rearranging equation (2) by vector identities, multiplying both terms by "one", and grouping appropriately,

gives the total angular momentum of the system of particles in terms ofmoment of inertiaandangular velocity,

Simplifications

Single particle

In the case of a single particle moving about the arbitrary origin,

and equations (**2) and (3**) for total angular momentum reduce to,

Fixed center of mass

For the case of the center of mass fixed in space with respect to the origin,

and equations (**2) and (3**) for total angular momentum reduce to,

Angular momentum (modern definition)

The 3-angular momentum as a bivector (plane element) and axial vector, of a particle of mass m with instantaneous 3-position x and 3-momentum p.

The 3-angular momentum as a bivector (plane element) and axial vector, of a particle of mass m with instantaneous 3-position x and 3-momentum p.

In modern (20th century) theoretical physics, angular momentum (not including any intrinsic angular momentum – see below) is described using a different formalism, instead of a classical pseudovector. In this formalism, angular momentum is the 2-form Noether charge associated with rotational invariance. As a result, angular momentum is not conserved for general curved spacetimes, unless it happens to be asymptotically rotationally invariant.

In classical mechanics, the angular momentum of a particle can be reinterpreted as a plane element:

in which the exterior product ∧ replaces the cross product × (these products have similar characteristics but are nonequivalent). This has the advantage of a clearer geometric interpretation as a plane element, defined from the x and p vectors, and the expression is true in any number of dimensions (two or higher). In Cartesian coordinates:

or more compactly in index notation:

The angular velocity can also be defined as an antisymmetric second order tensor, with components ωij. The relation between the two antisymmetric tensors is given by the moment of inertia which must now be a fourth order tensor:[26]

Again, this equation in L and ω as tensors is true in any number of dimensions. This equation also appears in the geometric algebra formalism, in which L and ω are bivectors, and the moment of inertia is a mapping between them.

In relativistic mechanics, the relativistic angular momentum of a particle is expressed as an antisymmetric tensor of second order:

in the language of four-vectors, namely the four position X and the four momentum P, and absorbs the above L together with the motion of the centre of mass of the particle.

In each of the above cases, for a system of particles, the total angular momentum is just the sum of the individual particle angular momenta, and the centre of mass is for the system.

In quantum mechanics

Angular momentum in quantum mechanics differs in many profound respects from angular momentum in classical mechanics. In relativistic quantum mechanics, it differs even more, in which the above relativistic definition becomes a tensorial operator.

Spin, orbital, and total angular momentum

Angular momenta of a classical object.Left: "spin" angular momentum S is really orbital angular momentum of the object at every point.Right: extrinsic orbital angular momentum L about an axis.Top: the moment of inertia tensor I and angular velocity ω (L is not always parallel to ω).Bottom: momentum p and its radial position r from the axis. The total angular momentum (spin plus orbital) is J. For a quantum particle the interpretations are different; particle spin does not have the above interpretation.

Angular momenta of a classical object.Left: "spin" angular momentum S is really orbital angular momentum of the object at every point.Right: extrinsic orbital angular momentum L about an axis.Top: the moment of inertia tensor I and angular velocity ω (L is not always parallel to ω).[27]Bottom: momentum p and its radial position r from the axis. The total angular momentum (spin plus orbital) is J. For a quantum particle the interpretations are different; particle spin does not have the above interpretation.

The classical definition of angular momentum ascan be carried over to quantum mechanics, by reinterpreting r as the quantumposition operatorand p as the quantummomentum operator. L is then anoperator, specifically called the *orbital angular momentum operator*. The components of the angular momentum operator satisfy the commutation relations of the Lie algebra so(3). Indeed, these operators are precisely the infinitesimal action of the rotation group on the quantum Hilbert space.[28] (See also the discussion below of the angular momentum operators as the generators of rotations.)

However, in quantum physics, there is another type of angular momentum, called spin angular momentum, represented by the spin operator S. Almost all elementary particles have spin. Spin is often depicted as a particle literally spinning around an axis, but this is a misleading and inaccurate picture: spin is an intrinsic property of a particle, unrelated to any sort of motion in space and fundamentally different from orbital angular momentum. All elementary particles have a characteristic spin, for example electrons have "spin 1/2" (this actually means "spin ħ/2") while photons have "spin 1" (this actually means "spin ħ").

Finally, there is total angular momentum J, which combines both the spin and orbital angular momentum of all particles and fields. (For one particle, J = L + S.) Conservation of angular momentum applies to J, but not to L or S; for example, the spin–orbit interaction allows angular momentum to transfer back and forth between L and S, with the total remaining constant. Electrons and photons need not have integer-based values for total angular momentum, but can also have fractional values.[29]

Quantization

In this standing wave on a circular string, the circle is broken into exactly 8 wavelengths. A standing wave like this can have 0,1,2, or any integer number of wavelengths around the circle, but it cannot have a non-integer number of wavelengths like 8.3. In quantum mechanics, angular momentum is quantized for a similar reason.

In this standing wave on a circular string, the circle is broken into exactly 8 wavelengths. A standing wave like this can have 0,1,2, or any integer number of wavelengths around the circle, but it cannot have a non-integer number of wavelengths like 8.3. In quantum mechanics, angular momentum is quantized for a similar reason.

Inquantum mechanics, angular momentum isquantized– that is, it cannot vary continuously, but only in "quantum leaps" between certain allowed values. For any system, the following restrictions on measurement results apply, whereis thereduced Planck constantandis anyEuclidean vectorsuch as x, y, or z:
If you measure…The result can be...
or
, where
or, where

(There are additional restrictions as well, see angular momentum operator for details.)

Thereduced Planck constantis tiny by everyday standards, about 10−34J s, and therefore this quantization does not noticeably affect the angular momentum of macroscopic objects. However, it is very important in the microscopic world. For example, the structure ofelectron shellsand subshells in chemistry is significantly affected by the quantization of angular momentum.

Quantization of angular momentum was first postulated by Niels Bohr in his Bohr model of the atom and was later predicted by Erwin Schrödinger in his Schrödinger equation.

Uncertainty

In the definition, six operators are involved: Theposition operators,,, and themomentum operators,,. However, theHeisenberg uncertainty principletells us that it is not possible for all six of these quantities to be known simultaneously with arbitrary precision. Therefore, there are limits to what can be known or measured about a particle's angular momentum. It turns out that the best that one can do is to simultaneously measure both the angular momentum vector'smagnitudeand its component along one axis.
The uncertainty is closely related to the fact that different components of anangular momentum operatordo notcommute, for example. (For the precisecommutation relations, seeangular momentum operator.)

Total angular momentum as generator of rotations

As mentioned above, orbital angular momentum L is defined as in classical mechanics:, but total angular momentum J is defined in a different, more basic way: J is defined as the "generator of rotations".[30] More specifically, J is defined so that the operator
is therotation operatorthat takes any system and rotates it by angleabout the axis. (The "exp" in the formula refers tooperator exponential) To put this the other way around, whatever our quantum Hilbert space is, we expect that therotation group SO(3)will act on it. There is then an associated action of the Lie algebra so(3) of SO(3); the operators describing the action of so(3) on our Hilbert space are the (total) angular momentum operators.

The relationship between the angular momentum operator and the rotation operators is the same as the relationship between Lie algebras and Lie groups in mathematics. The close relationship between angular momentum and rotations is reflected in Noether's theorem that proves that angular momentum is conserved whenever the laws of physics are rotationally invariant.

In electrodynamics

When describing the motion of a charged particle in an electromagnetic field, the canonical momentum P (derived from the Lagrangian for this system) is not gauge invariant. As a consequence, the canonical angular momentum L = r × P is not gauge invariant either. Instead, the momentum that is physical, the so-called kinetic momentum (used throughout this article), is (in SI units)

where e is the electric charge of the particle and A the magnetic vector potential of the electromagnetic field. The gauge-invariant angular momentum, that is kinetic angular momentum, is given by

The interplay with quantum mechanics is discussed further in the article on canonical commutation relations.

In optics

In classical Maxwell electrodynamics the Poynting vector is a linear momentum density of electromagnetic field.[31]

The angular momentum density vectoris given by a vector product as in classical mechanics:[32]
The above identities are valid locally, i.e. in each space pointin a given moment.

History

Newton, in the Principia, hinted at angular momentum in his examples of the First Law of Motion,

A top, whose parts by their cohesion are perpetually drawn aside from rectilinear motions, does not cease its rotation, otherwise than as it is retarded by the air. The greater bodies of the planets and comets, meeting with less resistance in more free spaces, preserve their motions both progressive and circular for a much longer time.[33]

He did not further investigate angular momentum directly in the Principia,

From such kind of reflexions also sometimes arise the circular motions of bodies about their own centres. But these are cases which I do not consider in what follows; and it would be too tedious to demonstrate every particular that relates to this subject.[34]

However, his geometric proof of the law of areas is an outstanding example of Newton's genius, and indirectly proves angular momentum conservation in the case of a central force.

The Law of Areas

Newton's derivation

Newton's derivation of the area law using geometric means.

Newton's derivation of the area law using geometric means.

As a planet orbits the Sun, the line between the Sun and the planet sweeps out equal areas in equal intervals of time. This had been known since Kepler expounded his second law of planetary motion. Newton derived a unique geometric proof, and went on to show that the attractive force of the Sun's gravity was the cause of all of Kepler's laws.

During the first interval of time, an object is in motion from point A to point B. Undisturbed, it would continue to point c during the second interval. When the object arrives at B, it receives an impulse directed toward point S. The impulse gives it a small added velocity toward S, such that if this were its only velocity, it would move from B to V during the second interval. By the rules of velocity composition, these two velocities add, and point C is found by construction of parallelogram BcCV. Thus the object's path is deflected by the impulse so that it arrives at point C at the end of the second interval. Because the triangles SBc and SBC have the same base SB and the same height Bc or VC, they have the same area. By symmetry, triangle SBc also has the same area as triangle SAB, therefore the object has swept out equal areas SAB and SBC in equal times.

At point C, the object receives another impulse toward S, again deflecting its path during the third interval from d to D. Thus it continues to E and beyond, the triangles SAB, SBc, SBC, SCd, SCD, SDe, SDE all having the same area. Allowing the time intervals to become ever smaller, the path ABCDE approaches indefinitely close to a continuous curve.

Note that because this derivation is geometric, and no specific force is applied, it proves a more general law than Kepler's second law of planetary motion. It shows that the Law of Areas applies to any central force, attractive or repulsive, continuous or non-continuous, or zero.

Conservation of angular momentum in the Law of Areas

The proportionality of angular momentum to the area swept out by a moving object can be understood by realizing that the bases of the triangles, that is, the lines from S to the object, are equivalent to the radius r, and that the heights of the triangles are proportional to the perpendicular component of velocity v⊥. Hence, if the area swept per unit time is constant, then by the triangular area formula 1/2(base)(height), the product (base)(height) and therefore the product rv⊥ are constant: if r and the base length are decreased, v⊥ and height must increase proportionally. Mass is constant, therefore angular momentum rmv⊥ is conserved by this exchange of distance and velocity.

In the case of triangle SBC, area is equal to 1/2(SB)(VC). Wherever C is eventually located due to the impulse applied at B, the product (SB)(VC), and therefore rmv⊥ remain constant. Similarly so for each of the triangles.

After Newton

Leonhard Euler, Daniel Bernoulli, and Patrick d'Arcy all understood angular momentum in terms of conservation of areal velocity, a result of their analysis of Kepler's second law of planetary motion. It is unlikely that they realized the implications for ordinary rotating matter.[35]

In 1736 Euler, like Newton, touched on some of the equations of angular momentum in his Mechanica without further developing them.[36]

Bernoulli wrote in a 1744 letter of a "moment of rotational motion", possibly the first conception of angular momentum as we now understand it.[37]

In 1799, Pierre-Simon Laplace first realized that a fixed plane was associated with rotation — his invariable plane.

Louis Poinsot in 1803 began representing rotations as a line segment perpendicular to the rotation, and elaborated on the "conservation of moments".

In 1852 Léon Foucault used a gyroscope in an experiment to display the Earth's rotation.

William J. M. Rankine's 1858 Manual of Applied Mechanics defined angular momentum in the modern sense for the first time:

...a line whose length is proportional to the magnitude of the angular momentum, and whose direction is perpendicular to the plane of motion of the body and of the fixed point, and such, that when the motion of the body is viewed from the extremity of the line, the radius-vector of the body seems to have right-handed rotation.

In an 1872 edition of the same book, Rankine stated that "The term angular momentum was introduced by Mr. Hayward,"[38] probably referring to R.B. Hayward's article On a Direct Method of estimating Velocities, Accelerations, and all similar Quantities with respect to Axes moveable in any manner in Space with Applications,[39] which was introduced in 1856, and published in 1864. Rankine was mistaken, as numerous publications feature the term starting in the late 18th to early 19th centuries.[40] However, Hayward's article apparently was the first use of the term and the concept seen by much of the English-speaking world. Before this, angular momentum was typically referred to as "momentum of rotation" in English.[41]

See also

  • Absolute angular momentum

  • Angular momentum coupling

  • Angular momentum of light

  • Angular momentum diagrams (quantum mechanics)

  • Clebsch–Gordan coefficients

  • Linear-rotational analogs

  • Orders of magnitude (angular momentum)

  • Pauli–Lubanski pseudovector

  • Relative angular momentum

  • Relativistic angular momentum

  • Rigid rotor

  • Rotational energy

  • Specific relative angular momentum

  • Yrast

References

[1]
Citation Linkbooks.google.comde Podesta, Michael (2002). Understanding the Properties of Matter (2nd, illustrated, revised ed.). CRC Press. p. 29. ISBN 978-0-415-25788-6. Extract of page 29
Sep 24, 2019, 3:52 AM
[2]
Citation Linkbooks.google.comWilson, E. B. (1915). Linear Momentum, Kinetic Energy and Angular Momentum. The American Mathematical Monthly. XXII. Ginn and Co., Boston, in cooperation with University of Chicago, et al. p. 190 – via Google books.
Sep 24, 2019, 3:52 AM
[3]
Citation Linkbooks.google.comWorthington, Arthur M. (1906). Dynamics of Rotation. Longmans, Green and Co., London. p. 21 – via Google books.
Sep 24, 2019, 3:52 AM
[4]
Citation Linkopenlibrary.orgTaylor, John R. (2005). Classical Mechanics. University Science Books, Mill Valley, CA. p. 90. ISBN 978-1-891389-22-1.
Sep 24, 2019, 3:52 AM
[5]
Citation Linkbooks.google.comDadourian, H. M. (1913). Analytical Mechanics for Students of Physics and Engineering. D. Van Nostrand Company, New York. p. 266 – via Google books.
Sep 24, 2019, 3:52 AM
[6]
Citation Linkbooks.google.comWatson, W. (1912). General Physics. Longmans, Green and Co., New York. p. 33 – via Google books.
Sep 24, 2019, 3:52 AM
[7]
Citation Linkbooks.google.comBarker, George F. (1893). Physics: Advanced Course (4th ed.). Henry Holt and Company, New York. p. 66 – via Google Books.
Sep 24, 2019, 3:52 AM
[8]
Citation Linkbooks.google.comBarker, George F. (1893). Physics: Advanced Course (4th ed.). Henry Holt and Company, New York. pp. 67–68 – via Google Books.
Sep 24, 2019, 3:52 AM
[9]
Citation Linkopenlibrary.orgOberg, Erik; et al. (2000). Machinery's Handbook (26th ed.). Industrial Press, Inc., New York. p. 143. ISBN 978-0-8311-2625-4.
Sep 24, 2019, 3:52 AM
[10]
Citation Linkbooks.google.comWatson, W. (1912). General Physics. Longmans, Green and Co., New York. p. 34 – via Google books.
Sep 24, 2019, 3:52 AM
[11]
Citation Linkbooks.google.comKent, William (1916). The Mechanical Engineers' Pocket Book (9th ed.). John Wiley and Sons, Inc., New York. p. 517 – via Google books.
Sep 24, 2019, 3:52 AM
[12]
Citation Linkopenlibrary.orgOberg, Erik; et al. (2000). Machinery's Handbook (26th ed.). Industrial Press, Inc., New York. p. 146. ISBN 978-0-8311-2625-4.
Sep 24, 2019, 3:52 AM
[13]
Citation Linkopenlibrary.orgOberg, Erik; et al. (2000). Machinery's Handbook (26th ed.). Industrial Press, Inc., New York. pp. 161–162. ISBN 978-0-8311-2625-4.
Sep 24, 2019, 3:52 AM
[14]
Citation Linkbooks.google.comKent, William (1916). The Mechanical Engineers' Pocket Book (9th ed.). John Wiley and Sons, Inc., New York. p. 527 – via Google books.
Sep 24, 2019, 3:52 AM
[15]
Citation Linkopenlibrary.orgBattin, Richard H. (1999). An Introduction to the Mathematics and Methods of Astrodynamics, Revised Edition. American Institute of Aeronautics and Astronautics, Inc. ISBN 978-1-56347-342-5., p. 97
Sep 24, 2019, 3:52 AM
[16]
Citation Linkbooks.google.comRankine, W. J. M. (1872). A Manual of Applied Mechanics (6th ed.). Charles Griffin and Company, London. p. 507 – via Google books.
Sep 24, 2019, 3:52 AM
[17]
Citation Linkbooks.google.comCrew, Henry (1908). The Principles of Mechanics: For Students of Physics and Engineering. Longmans, Green, and Company, New York. p. 88 – via Google books.
Sep 24, 2019, 3:52 AM
[18]
Citation Linkbooks.google.comWorthington, Arthur M. (1906). Dynamics of Rotation. Longmans, Green and Co., London. p. 82 – via Google books.
Sep 24, 2019, 3:52 AM
[19]
Citation Linkbooks.google.comWorthington, Arthur M. (1906). Dynamics of Rotation. Longmans, Green and Co., London. p. 11 – via Google books.
Sep 24, 2019, 3:52 AM
[20]
Citation Link//doi.org/10.1098%2Frsta.1984.0082Stephenson, F. R.; Morrison, L. V.; Whitrow, G. J. (1984). "Long-term changes in the rotation of the earth – 700 B.C. to A.D. 1980". Philosophical Transactions of the Royal Society. 313 (1524): 67. Bibcode:1984RSPTA.313...47S. doi:10.1098/rsta.1984.0082. +2.40 ms/century divided by 36525 days.
Sep 24, 2019, 3:52 AM