Bertrand's theorem: Difference between revisions
(29 intermediate revisions by 21 users not shown) | |||
Line 1: | Line 1: | ||
{{Short description|Physics theorem}} |
|||
[[File:Bertrand.jpg|thumb|right|200px|Joseph Bertrand]] |
[[File:Bertrand.jpg|thumb|right|200px|Joseph Bertrand]] |
||
In [[ |
In [[classical mechanics]], '''Bertrand's theorem''' states that among [[central-force]] [[potential]]s with bound orbits, there are only two types of [[central-force]] (radial) [[scalar potential]]s with the property that all bound orbits are also [[orbit (dynamics)|closed orbits]].<ref>{{cite journal | author = Bertrand J | author-link = Joseph Louis François Bertrand | year = 1873 | title = Théorème relatif au mouvement d'un point attiré vers un centre fixe | journal = C. R. Acad. Sci. | volume = 77 | pages = 849–853}}</ref><ref name="Johnson2010">{{cite book |last=Johnson |first=Porter Wear |title=Classical Mechanics With Applications |url=https://books.google.com/books?id=SKLjx5Qfst8C&pg=PA149 |accessdate=2 December 2012 |date=2010-02-24 |publisher=World Scientific |isbn=9789814304153 |pages=149–}}</ref> |
||
⚫ | |||
⚫ | |||
: <math>V(\mathbf{r}) = -\frac{k}{r},</math> |
|||
: <math>V(r) = -\frac{k}{r}</math> with force <math>f(r) = -\frac{dV}{dr} = -\frac{k}{r^2}</math>. |
|||
⚫ | |||
⚫ | |||
: <math>V(\mathbf{r}) = \frac{1}{2} kr^2.</math> |
|||
: <math>V(r) = \frac{1}{2} kr^2</math> with force <math>f(r) = -\frac{dV}{dr} = -kr</math>. |
|||
==Description== |
|||
The theorem is named after its discoverer, [[Joseph Bertrand]]. |
|||
==Derivation== |
|||
[[File:Closed orbits.gif|300px|thumb|right|Small changes in the power of the force scaling with distance will lead to significantly different kind of orbits.]] |
|||
All attractive [[central force]]s can produce [[circle|circular]] orbits, which are naturally [[orbit (dynamics)|closed orbits]]. The only requirement is that the central force exactly equals the [[centripetal force]], which determines the required angular velocity for a given circular radius. Non-central forces (i.e., those that depend on the angular variables as well as the radius) are ignored here, since they do not produce circular orbits in general. |
All attractive [[central force]]s can produce [[circle|circular]] orbits, which are naturally [[orbit (dynamics)|closed orbits]]. The only requirement is that the central force exactly equals the [[centripetal force]], which determines the required angular velocity for a given circular radius. Non-central forces (i.e., those that depend on the angular variables as well as the radius) are ignored here, since they do not produce circular orbits in general. |
||
The equation of motion for the radius |
The equation of motion for the radius <math>r</math> of a particle of mass <math>m</math> moving in a [[central force|central potential]] <math>V(r)</math> is given by [[Euler–Lagrange equation|motion equations]] |
||
: <math>m\frac{d^2 r}{dt^2} - mr \omega^2 = m\frac{d^2 r}{dt^2} - \frac{L^2}{mr^3} = -\frac{dV}{dr},</math> |
: <math>m\frac{d^2 r}{dt^2} - mr \omega^2 = m\frac{d^2 r}{dt^2} - \frac{L^2}{mr^3} = -\frac{dV}{dr},</math> |
||
where <math>\omega \equiv \frac{d\theta}{dt}</math>, and the [[angular momentum]] |
where <math>\omega \equiv \frac{d\theta}{dt}</math>, and the [[angular momentum]] <math>L = mr^2\omega</math> is conserved. For illustration, the first term on the left is zero for circular orbits, and the applied inwards force <math>\frac{dV}{dr}</math> equals the [[centripetal force|centripetal force requirement]] <math> mr^2\omega</math>, as expected. |
||
The definition of [[angular momentum]] allows a change of independent variable from |
The definition of [[angular momentum]] allows a change of independent variable from <math>t</math> to <math>\theta</math>: |
||
: <math>\frac{d}{dt} = \frac{L}{mr^2} \frac{d}{d\theta},</math> |
: <math>\frac{d}{dt} = \frac{L}{mr^2} \frac{d}{d\theta},</math> |
||
Line 30: | Line 35: | ||
: <math>\frac{d^2 u}{d\theta^2} + u = -\frac{m}{L^2} \frac{d}{du} V\left(\frac{1}{u}\right).</math> |
: <math>\frac{d^2 u}{d\theta^2} + u = -\frac{m}{L^2} \frac{d}{du} V\left(\frac{1}{u}\right).</math> |
||
As noted above, all [[central force]]s can produce [[orbit (dynamics)|circular orbits]] given an appropriate initial velocity. However, if some radial velocity is introduced, these orbits need not be stable (i.e., remain in orbit indefinitely) nor closed (repeatedly returning to exactly the same path). Here we show that stable, exactly closed orbits |
As noted above, all [[central force]]s can produce [[orbit (dynamics)|circular orbits]] given an appropriate initial velocity. However, if some radial velocity is introduced, these orbits need not be stable (i.e., remain in orbit indefinitely) nor closed (repeatedly returning to exactly the same path). Here we show that a necessary condition for stable, exactly closed non-circular orbits is an inverse-square force or radial harmonic oscillator potential. In the following sections, we show that those two force laws produce [[orbit (dynamics)|stable, exactly closed orbits]]. |
||
Define |
Define <math>J(u)</math> as |
||
: <math> |
: <math> |
||
Line 40: | Line 45: | ||
</math> |
</math> |
||
where |
where <math>f</math> represents the radial force. The criterion for perfectly [[circle|circular]] motion at a radius <math>r_0</math> is that the first term on the left be zero: |
||
{{NumBlk|:|<math>u_0 = J(u_0) = -\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right),</math>|{{EquationRef|1}}}} |
{{NumBlk|:|<math>u_0 = J(u_0) = -\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right),</math>|{{EquationRef|1}}}} |
||
Line 46: | Line 51: | ||
where <math>u_0 \equiv 1/r_0</math>. |
where <math>u_0 \equiv 1/r_0</math>. |
||
The next step is to consider the equation for |
The next step is to consider the equation for <math>u</math> under ''[[perturbation theory|small perturbations]]'' <math>\eta \equiv u - u_0</math> from perfectly circular orbits. On the right, the <math>J</math> function can be expanded in a standard [[Taylor series]]: |
||
: <math>J(u) \approx J(u_0) + \eta J'(u_0) + \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_0) + \cdots</math> |
: <math>J(u) \approx J(u_0) + \eta J'(u_0) + \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_0) + \cdots</math> |
||
Substituting this expansion into the equation for |
Substituting this expansion into the equation for <math>u</math> and subtracting the constant terms yields |
||
: <math>\frac{d^2 \eta}{d\theta^2} + \eta = \eta J'(u_0) + \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_0) + \cdots,</math> |
: <math>\frac{d^2 \eta}{d\theta^2} + \eta = \eta J'(u_0) + \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_0) + \cdots,</math> |
||
Line 58: | Line 63: | ||
{{NumBlk|:|<math>\frac{d^2 \eta}{d\theta^2} + \beta^2 \eta = \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_{0}) + \cdots,</math>|{{EquationRef|2}}}} |
{{NumBlk|:|<math>\frac{d^2 \eta}{d\theta^2} + \beta^2 \eta = \frac{1}{2} \eta^2 J''(u_0) + \frac{1}{6} \eta^3 J'''(u_{0}) + \cdots,</math>|{{EquationRef|2}}}} |
||
where <math>\beta^2 \equiv 1 - J'(u_0)</math> is a constant. |
where <math>\beta^2 \equiv 1 - J'(u_0)</math> is a constant. <math>\beta^2</math> must be non-negative; otherwise, the radius of the orbit would vary exponentially away from its initial radius. (The solution <math>\beta=0</math> corresponds to a perfectly circular orbit.) If the right side may be neglected (i.e., for small perturbations), the solutions are |
||
: <math>\eta(\theta) = h_1 \cos(\beta\theta),</math> |
: <math>\eta(\theta) = h_1 \cos(\beta\theta),</math> |
||
where the amplitude |
where the amplitude <math>h_1</math> is a constant of integration. For the orbits to be closed, <math>\beta</math> must be a [[rational number]]. What's more, it must be the ''same'' rational number for all radii, since <math>\beta</math> cannot change continuously; the [[rational number]]s are [[totally disconnected space|totally disconnected]] from one another. Using the definition of <math>J</math> along with equation ({{EquationNote|1}}), |
||
: <math> |
: <math> |
||
J'(u_0) = |
J'(u_0) = |
||
\frac{2}{u_0} \left[\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right)\right] - \left[\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right)\right]\frac{1}{f\left(\frac{1}{u_0}\right)} \frac{ |
\frac{2}{u_0} \left[\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right)\right] - \left[\frac{m}{L^2 u_0^2} f\left(\frac{1}{u_0}\right)\right]\frac{1}{f\left(\frac{1}{u_0}\right)} \frac{d}{du_0}f\left(\frac1{u_0}\right) = |
||
-2 + \frac{u_0}{f\left(\frac{1}{u_0}\right)} \frac{ |
-2 + \frac{u_0}{f\left(\frac{1}{u_0}\right)} \frac{d}{du_0}f\left(\frac1{u_0}\right) = |
||
1 - \beta^2 |
1 - \beta^2. |
||
</math> |
</math> |
||
Since this must hold for any value of <math>u_0</math>, |
|||
: <math>\frac{df}{dr} = (\beta^2 - 3) \frac{f}{r},</math> |
: <math>\frac{df}{dr} = (\beta^2 - 3) \frac{f}{r},</math> |
||
Line 79: | Line 84: | ||
: <math>f(r) = -\frac{k}{r^{3-\beta^2}}.</math> |
: <math>f(r) = -\frac{k}{r^{3-\beta^2}}.</math> |
||
Hence, |
Hence, <math>J</math> must have the general form |
||
{{NumBlk|:|<math>J(u) = \frac{mk}{L^2} u^{1-\beta^2}.</math>|{{EquationRef|3}}}} |
{{NumBlk|:|<math>J(u) = \frac{mk}{L^2} u^{1-\beta^2}.</math>|{{EquationRef|3}}}} |
||
For more general deviations from circularity (i.e., when we cannot neglect the higher-order terms in the Taylor expansion of |
For more general deviations from circularity (i.e., when we cannot neglect the higher-order terms in the Taylor expansion of <math>J</math> ), <math>\eta</math> may be expanded in a Fourier series, e.g., |
||
: <math>\eta(\theta) = h_0 + h_1 \cos \beta \theta + h_2 \cos 2\beta \theta + h_3 \cos 3\beta \theta + \cdots</math> |
: <math>\eta(\theta) = h_0 + h_1 \cos \beta \theta + h_2 \cos 2\beta \theta + h_3 \cos 3\beta \theta + \cdots</math> |
||
We substitute this into equation ({{EquationNote|2}}) and equate the coefficients belonging to the same frequency, keeping only the lowest-order terms. As we show below, |
We substitute this into equation ({{EquationNote|2}}) and equate the coefficients belonging to the same frequency, keeping only the lowest-order terms. As we show below, <math>h_0</math> and <math>h_2</math> are smaller than <math>h_1</math>, being of order <math>h_1^2</math>. <math>h_3</math>, and all further coefficients, are at least of order <math>h_1^3</math>. This makes sense, since <math>h_0, h_2, h_3, \ldots</math> must all vanish faster than <math>h_1</math> as a circular orbit is approached. |
||
: <math>h_0 = h_1^2 \frac{J''(u_0)}{4\beta^2},</math> |
: <math>h_0 = h_1^2 \frac{J''(u_0)}{4\beta^2},</math> |
||
Line 93: | Line 98: | ||
: <math>h_3 = -\frac{1}{8\beta^2} \left[ h_1 h_2 \frac{J''(u_0)}{2} + h_1^3 \frac{J'''(u_0)}{24} \right].</math> |
: <math>h_3 = -\frac{1}{8\beta^2} \left[ h_1 h_2 \frac{J''(u_0)}{2} + h_1^3 \frac{J'''(u_0)}{24} \right].</math> |
||
From the cos |
From the <math>\cos\beta\theta</math> term, we get |
||
: <math> |
: <math> |
||
Line 101: | Line 106: | ||
</math> |
</math> |
||
where in the last step we substituted in the values of |
where in the last step we substituted in the values of <math>h_0</math> and <math>h_2</math>. |
||
Using equations ({{EquationNote|3}}) and ({{EquationNote|1}}), we can calculate the second and third derivatives of |
Using equations ({{EquationNote|3}}) and ({{EquationNote|1}}), we can calculate the second and third derivatives of <math>J</math> evaluated at <math>u_0</math>: |
||
: <math>J''(u_0) = -\frac{\beta^2 (1 - \beta^2)}{u_0},</math> |
: <math>J''(u_0) = -\frac{\beta^2 (1 - \beta^2)}{u_0},</math> |
||
Line 112: | Line 117: | ||
: <math>\beta^2 (1 - \beta^2) (4 - \beta^2) = 0.</math> |
: <math>\beta^2 (1 - \beta^2) (4 - \beta^2) = 0.</math> |
||
Hence, the only [[potential]]s that can produce stable closed non-circular orbits are the inverse-square force law ( |
Hence, the only [[potential]]s that can produce stable closed non-circular orbits are the inverse-square force law (<math>\beta=1</math>) and the radial harmonic-oscillator potential (<math>\beta=2</math>). The solution <math>\beta=0</math> corresponds to perfectly circular orbits, as noted above. |
||
==Classical field potentials== |
==Classical field potentials== |
||
For an inverse-square force law such as the [[gravity|gravitational]] or [[electrostatics|electrostatic potential]], the [[potential]] can be written |
For an inverse-square force law such as the [[gravity|gravitational]] or [[electrostatics|electrostatic potential]], the [[potential]] can be written |
||
Line 129: | Line 134: | ||
where ''e'' (the ''eccentricity''), and θ<sub>0</sub> (the ''phase offset'') are constants of integration. |
where ''e'' (the ''eccentricity''), and θ<sub>0</sub> (the ''phase offset'') are constants of integration. |
||
This is the general formula for a [[conic section]] that has one focus at the origin; ''e'' = 0 corresponds to a [[circle]], ''e'' < 1 corresponds to an ellipse, ''e'' = 1 corresponds to a [[parabola]], and ''e'' > 1 corresponds to a [[hyperbola]]. The eccentricity ''e'' is related to the total [[energy]] ''E'' (see [[Laplace–Runge–Lenz vector]]): |
This is the general formula for a [[conic section]] that has one focus at the origin; ''e'' = 0 corresponds to a [[circle]], 0 < ''e'' < 1 corresponds to an ellipse, ''e'' = 1 corresponds to a [[parabola]], and ''e'' > 1 corresponds to a [[hyperbola]]. The eccentricity ''e'' is related to the total [[energy]] ''E'' (see [[Laplace–Runge–Lenz vector]]): |
||
: <math>e = \sqrt{1 + \frac{2EL^2}{k^2 m}}.</math> |
: <math>e = \sqrt{1 + \frac{2EL^2}{k^2 m}}.</math> |
||
Line 152: | Line 157: | ||
: <math>z = A_z \cos(\omega_0 t + \phi_z),</math> |
: <math>z = A_z \cos(\omega_0 t + \phi_z),</math> |
||
where the positive constants ''A<sub>x</sub>'', ''A<sub>y</sub>'' and ''A<sub>z</sub>'' represent the ''amplitudes'' of the oscillations, and the angles φ<sub>''x''</sub>, φ<sub>''y''</sub> and φ<sub>''z''</sub> represent their ''phases''. The resulting orbit '''r'''(''t'') = [''x''(''t''), ''y''(''y''), ''z''(''t'')] is closed because it repeats exactly after |
where the positive constants ''A<sub>x</sub>'', ''A<sub>y</sub>'' and ''A<sub>z</sub>'' represent the ''amplitudes'' of the oscillations, and the angles φ<sub>''x''</sub>, φ<sub>''y''</sub> and φ<sub>''z''</sub> represent their ''phases''. The resulting orbit '''r'''(''t'') = [''x''(''t''), ''y''(''y''), ''z''(''t'')] is closed because it repeats exactly after one period |
||
: <math>T \equiv \frac{2\pi}{\omega_0}.</math> |
: <math>T \equiv \frac{2\pi}{\omega_0}.</math> |
||
Line 159: | Line 164: | ||
==References== |
==References== |
||
{{ |
{{Reflist|1}} |
||
==Further reading== |
==Further reading== |
||
Line 180: | Line 185: | ||
| arxiv=0704.2396 |
| arxiv=0704.2396 |
||
| bibcode=2007arXiv0704.2396S |
| bibcode=2007arXiv0704.2396S |
||
}} |
|||
*{{cite journal |
|||
| title = A Comprehensive Proof of Bertrand's Theorem |
|||
| last1 = Leenheer | first1 = Patrick De |
|||
| last2 = Musgrove | first2 = John |
|||
| last3 = Schimleck | first3 = Tyler |
|||
| date=2023 |
|||
| journal=SIAM Review |
|||
| url=https://epubs.siam.org/doi/10.1137/21M1436658 |
|||
| volume=65 | issue=2 | pages=563–588 |
|||
| doi=10.1137/21M1436658 |
|||
| s2cid = 258585586 | issn=0036-1445 |
|||
}} |
}} |
||
[[Category:Classical mechanics]] |
[[Category:Classical mechanics]] |
||
⚫ | |||
[[Category:Orbits]] |
[[Category:Orbits]] |
||
⚫ |
Latest revision as of 15:03, 16 December 2024
In classical mechanics, Bertrand's theorem states that among central-force potentials with bound orbits, there are only two types of central-force (radial) scalar potentials with the property that all bound orbits are also closed orbits.[1][2]
The first such potential is an inverse-square central force such as the gravitational or electrostatic potential:
- with force .
The second is the radial harmonic oscillator potential:
- with force .
The theorem is named after its discoverer, Joseph Bertrand.
Derivation
[edit]All attractive central forces can produce circular orbits, which are naturally closed orbits. The only requirement is that the central force exactly equals the centripetal force, which determines the required angular velocity for a given circular radius. Non-central forces (i.e., those that depend on the angular variables as well as the radius) are ignored here, since they do not produce circular orbits in general.
The equation of motion for the radius of a particle of mass moving in a central potential is given by motion equations
where , and the angular momentum is conserved. For illustration, the first term on the left is zero for circular orbits, and the applied inwards force equals the centripetal force requirement , as expected.
The definition of angular momentum allows a change of independent variable from to :
giving the new equation of motion that is independent of time:
This equation becomes quasilinear on making the change of variables and multiplying both sides by (see also Binet equation):
As noted above, all central forces can produce circular orbits given an appropriate initial velocity. However, if some radial velocity is introduced, these orbits need not be stable (i.e., remain in orbit indefinitely) nor closed (repeatedly returning to exactly the same path). Here we show that a necessary condition for stable, exactly closed non-circular orbits is an inverse-square force or radial harmonic oscillator potential. In the following sections, we show that those two force laws produce stable, exactly closed orbits.
Define as
where represents the radial force. The criterion for perfectly circular motion at a radius is that the first term on the left be zero:
(1) |
where .
The next step is to consider the equation for under small perturbations from perfectly circular orbits. On the right, the function can be expanded in a standard Taylor series:
Substituting this expansion into the equation for and subtracting the constant terms yields
which can be written as
(2) |
where is a constant. must be non-negative; otherwise, the radius of the orbit would vary exponentially away from its initial radius. (The solution corresponds to a perfectly circular orbit.) If the right side may be neglected (i.e., for small perturbations), the solutions are
where the amplitude is a constant of integration. For the orbits to be closed, must be a rational number. What's more, it must be the same rational number for all radii, since cannot change continuously; the rational numbers are totally disconnected from one another. Using the definition of along with equation (1),
Since this must hold for any value of ,
which implies that the force must follow a power law
Hence, must have the general form
(3) |
For more general deviations from circularity (i.e., when we cannot neglect the higher-order terms in the Taylor expansion of ), may be expanded in a Fourier series, e.g.,
We substitute this into equation (2) and equate the coefficients belonging to the same frequency, keeping only the lowest-order terms. As we show below, and are smaller than , being of order . , and all further coefficients, are at least of order . This makes sense, since must all vanish faster than as a circular orbit is approached.
From the term, we get
where in the last step we substituted in the values of and .
Using equations (3) and (1), we can calculate the second and third derivatives of evaluated at :
Substituting these values into the last equation yields the main result of Bertrand's theorem:
Hence, the only potentials that can produce stable closed non-circular orbits are the inverse-square force law () and the radial harmonic-oscillator potential (). The solution corresponds to perfectly circular orbits, as noted above.
Classical field potentials
[edit]For an inverse-square force law such as the gravitational or electrostatic potential, the potential can be written
The orbit u(θ) can be derived from the general equation
whose solution is the constant plus a simple sinusoid:
where e (the eccentricity), and θ0 (the phase offset) are constants of integration.
This is the general formula for a conic section that has one focus at the origin; e = 0 corresponds to a circle, 0 < e < 1 corresponds to an ellipse, e = 1 corresponds to a parabola, and e > 1 corresponds to a hyperbola. The eccentricity e is related to the total energy E (see Laplace–Runge–Lenz vector):
Comparing these formulae shows that E < 0 corresponds to an ellipse, E = 0 corresponds to a parabola, and E > 0 corresponds to a hyperbola. In particular, for perfectly circular orbits.
Harmonic oscillator
[edit]To solve for the orbit under a radial harmonic-oscillator potential, it's easier to work in components r = (x, y, z). The potential can be written as
The equation of motion for a particle of mass m is given by three independent Euler equations:
where the constant must be positive (i.e., k > 0) to ensure bounded, closed orbits; otherwise, the particle will fly off to infinity. The solutions of these simple harmonic oscillator equations are all similar:
where the positive constants Ax, Ay and Az represent the amplitudes of the oscillations, and the angles φx, φy and φz represent their phases. The resulting orbit r(t) = [x(t), y(y), z(t)] is closed because it repeats exactly after one period
The system is also stable because small perturbations in the amplitudes and phases cause correspondingly small changes in the overall orbit.
References
[edit]- ^ Bertrand J (1873). "Théorème relatif au mouvement d'un point attiré vers un centre fixe". C. R. Acad. Sci. 77: 849–853.
- ^ Johnson, Porter Wear (2010-02-24). Classical Mechanics With Applications. World Scientific. pp. 149–. ISBN 9789814304153. Retrieved 2 December 2012.
Further reading
[edit]- Goldstein, H. (1980). Classical Mechanics (2nd ed.). Addison-Wesley. ISBN 978-0-201-02918-5.
- Santos, F. C.; Soares, V.; Tort, A. C. (2011). "An English translation of Bertrand's theorem". Latin American Journal of Physics Education. 5 (4): 694–696. arXiv:0704.2396. Bibcode:2007arXiv0704.2396S.
- Leenheer, Patrick De; Musgrove, John; Schimleck, Tyler (2023). "A Comprehensive Proof of Bertrand's Theorem". SIAM Review. 65 (2): 563–588. doi:10.1137/21M1436658. ISSN 0036-1445. S2CID 258585586.