Main Page: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
No edit summary
mNo edit summary
Line 1: Line 1:
The '''missionaries and cannibals problem''', and the closely related '''jealous husbands problem''', are classic [[river-crossing problems]].<ref name=a>[http://links.jstor.org/sici?sici=0025-5572%28198906%292%3A73%3A464%3C73%3A%22JHA%22M%3E2.0.CO%3B2-S "The Jealous Husbands" and "The Missionaries and Cannibals"], Ian Pressman and David Singmaster, ''The Mathematical Gazette'', '''73''', #464 (June 1989), pp. 73&ndash;81.</ref>  The missionaries and cannibals problem is a well-known [[toy problem]] in [[artificial intelligence]], where it was used by [[Saul Amarel]] as an example of problem representation.<ref>[http://www.cc.gatech.edu/~jimmyd/summaries/amarel1968-1.html On representations of problems of reasoning about actions], Saul Amarel, pp. 131&ndash;171, ''Machine Intelligence 3'', edited by Donald Michie, Amsterdam, London, New York: Elsevier/North-Holland, 1968.</ref><ref>p. 9, [http://dx.doi.org/10.1007/11829263_1 Searching in a Maze, in Search of Knowledge: Issues in Early Artificial Intelligence], Roberto Cordeschi, pp. 1&ndash;23, ''Reasoning, Action andj Interaction in AI Theories and Systems: essays dedicated to Luigia Carlucci Aiello'', edited by Oliviero Stock and Marco Schaerf, Lecture Notes in Computer Science #4155, Berlin/Heidelberg: Springer, 2006, ISBN 978-3-540-37901-0.</ref>
An '''adiabatic invariant''' is a property of a [[physical system]] that stays constant when changes occur slowly.


==The problem==
In [[thermodynamics]], an adiabatic process is a change that occurs without heat flow, and slowly compared to the time to reach equilibrium. In an adiabatic process, the system is in equilibrium at all stages. Under these conditions, the [[entropy]] is constant.
In the missionaries and cannibals problem, three missionaries and three cannibals must cross a river using a boat which can carry at most two people, under the constraint that, for both banks, if there are missionaries present on the bank, they cannot be outnumbered by cannibals (if they were, the cannibals would eat the missionaries.)  The boat cannot cross the river by itself with no people on board.<ref name=a />


In the jealous husbands problem, the missionaries and cannibals become three married couples, with the constraint that no woman can be in the presence of another man unless her husband is also present. Under this constraint, there cannot be both women and men present on a bank with women outnumbering men, since if there were, some woman would be husbandless.  Therefore, upon changing men to missionaries and women to cannibals, any solution to the jealous husbands problem will also become a solution to the missionaries and cannibals problem.<ref name=a />
In [[mechanics]], an adiabatic change is a slow deformation of the [[Hamiltonian (quantum mechanics)|Hamiltonian]], where the [[logarithmic derivative|fractional rate of change]] of the energy is much slower than the orbital frequency. The area enclosed by the different motions in phase space are the ''adiabatic invariants''.


==Solving==
In [[quantum mechanics]], an adiabatic change is one that occurs at a rate much slower than the difference in frequency between energy eigenstates. In this case, the energy states of the system do not make transitions, so that the [[quantum number]] is an adiabatic invariant.
Amarel devised a system for solving the Missionaries and Cannibals problem whereby the current state is represented by a simple vector <a,b,c>. The vector's elements represent the number of missionaries on the wrong side, the number of cannibals on the wrong side, and the number of boats on the wrong side, respectively. Since the boat and all of the missionaries and cannibals start on the wrong side, the vector is initialized to <3,3,1>. Actions are represented using vector subtraction/addition to manipulate the state vector. For instance, if a lone cannibal crossed the river, the vector <0,1,1> would be subtracted from the state to yield <3,2,0>. The state would reflect that there are still three missionaries and two cannibals on the wrong side, and that the boat is now on the opposite bank. To fully solve the problem, a simple tree is formed with the initial state as the root. The five possible actions (<1,0,1>, <2,0,1>, <0,1,1>, <0,2,1>, and <1,1,1>) are then ''subtracted'' from the initial state, with the result forming children nodes of the root. Any node that has more cannibals than missionaries on either bank is in an invalid state, and is therefore removed from further consideration. The valid children nodes generated would be <3,2,0>, <3,1,0>, and <2,2,0>. For each of these remaining nodes, children nodes are generated by ''adding'' each of the possible action vectors. The algorithm continues alternating subtraction and addition for each level of the tree until a node is generated with the vector <0,0,0> as its value. This is the goal state, and the path from the root of the tree to this node represents a sequence of actions that solves the problem.


==Solution==
The [[old quantum theory]] was formulated by equating the quantum number of a system with its classical adiabatic invariant. This determined the form of the Bohr-Sommerfeld quantization rule: the quantum number is the area in phase space of the classical orbit.
The earliest solution known to the jealous husbands problem, using 11 one-way trips, is as follows.  The married couples are represented as ''<math>\alpha</math>'' (male) and ''a'' (female), ''<math>\beta</math>'' and ''b'', and ''<math>\gamma</math>'' and ''c''.<ref name=b>Jealous Husbands Crossing the River: A Problem from Alcuin to Tartaglia,
Raffaella Franci, pp. 289&ndash;306, ''From China to Paris: 2000 Years Transmission of Mathematical Ideas'', edited by Yvonne Dold-Samplonius, Joseph W. Dauben, Menso Folkerts, and Benno van Dalen, Stuttgart: Franz Steiner Verlag, 2002, ISBN 3-515-08223-9.</ref><sup>,&nbsp;p.&nbsp;291.</sup>


{| class="wikitable"
== Thermodynamics ==
! Trip number
! Starting bank
! Travel
! Ending bank
|-
| (start)
| <math>\alpha</math>a  <math>\beta</math>b <math>\gamma</math>c
|
|
|-
|1
|<math>\beta</math>b <math>\gamma</math>c
|<math>\alpha</math>a →
|
|-
|2
|<math>\beta</math>b <math>\gamma</math>c
|←<math>\alpha</math>
|a
|-
|3
|<math>\alpha</math> <math>\beta</math> <math>\gamma</math>
|bc →
|a
|-
|4
|<math>\alpha</math> <math>\beta</math> <math>\gamma</math>
| ← a
|b c
|-
|5
|<math>\alpha</math>a
| <math>\beta</math><math>\gamma</math> →
|b c
|-
|6
|<math>\alpha</math>a
|← <math>\beta</math>b
|<math>\gamma</math>c
|-
|7
|a b
|<math>\alpha</math><math>\beta</math> →
|<math>\gamma</math>c
|-
|8
|a b
|← c
|<math>\alpha</math> <math>\beta</math> <math>\gamma</math>
|-©
|9
|b
|a c →
|<math>\alpha</math> <math>\beta</math> <math>\gamma</math>
|-
|10
|b
|← <math>\beta</math>
|<math>\alpha</math>a <math>\gamma</math>c
|-
|11
|
|<math>\beta</math>b →
|<math>\alpha</math>a <math>\gamma</math>c
|-
|(finish)
|
|
|<math>\alpha</math>a <math>\beta</math>b <math>\gamma</math>c
|}


This is a shortest solution to the problem, but is not the only shortest solution.<ref name=b /><sup>,&nbsp;p.&nbsp;291.</sup>
In thermodynamics, adiabatic changes are those that do not increase the entropy. They occur slowly, and allow heat flow only between objects at the same temperature. For isolated systems, an adiabatic change allows no heat to flow in or out.


If however, only one man can get out of the boat at a time and husbands must be on the shore to count as with his wife as opposed to just being in the boat at the shore: move 5 to 6 is impossible, for as soon as ''<math>\gamma</math>'' has stepped out ''b'' on the shore won't be with her husband, despite him being just in the boat.
=== Adiabatic expansion of an ideal gas ===


As mentioned previously, this solution to the jealous husbands problem will become a solution to the missionaries and cannibals problem upon replacing men by missionaries and women by cannibals.  In this case we may neglect the individual identities of the missionaries and cannibals. The solution just given is still shortest, and is one of four shortest solutions.<ref>[http://portal.acm.org/citation.cfm?id=144045.144106 Cannibals and missionaries], Ruby Lim, pp. 135&ndash;142, ''Conference proceedings: APL '92, the International Conference on APL, 6–10 July 1992, St. Petersburg, Russia'', edited by Lynne C. Shaw, et al., New York: Association for Computing Machinery, 1992, ISBN 0-89791-477-5.</ref>
If a container with an [[ideal gas]] is expanded instantaneously, the temperature of the gas doesn't change at all, because none of the molecules slow down. The molecules keep their kinetic energy, but now the gas occupies a bigger volume. If the container expands slowly, however, so that the ideal gas pressure law holds at any time, gas molecules lose energy at the rate that they do work on the expanding wall. The amount of work they do is the pressure times the area of the wall times the outward displacement, which is the pressure times the change in the volume of the gas:
:<math>
dW = P dV = {N k_B T \over V} dV
</math>


If a woman in the boat at the shore (but not ''on'' the shore) counts as being by herself (i.e. not in the presence of any men on the shore), then this puzzle can be solved in 9 one-way trips:
If no heat enters the gas, the energy in the gas molecules is decreasing by the same amount. By definition, a gas is ideal when its temperature is only a function of the internal energy per particle, not the volume. So
:<math>
dT = {1 \over N C_v} dE
</math>


{| class="wikitable"
Where <math>C_{v}</math> is the specific heat at constant volume. When the change in energy is entirely due to work done on the wall, the change in temperature is given by:
! Trip number
:<math>
! Starting bank
N C_v dT = - dW = - {N{k_B}T \over V} dV
! Travel
</math>
! Ending bank
|-
| (start)
|<math>\alpha</math>a <math>\beta</math>b <math>\gamma</math>c
|
|
|-
|1
|<math>\beta</math>b <math>\gamma</math>c
|<math>\alpha</math>a →
|
|-
|2
|<math>\beta</math>b <math>\gamma</math>c
|← a
|<math>\alpha</math>
|-
|3
|<math>\beta</math> <math>\gamma</math>c
|ab →
|<math>\alpha</math>
|-
|4
|<math>\beta</math> <math>\gamma</math>c
| ← b
|<math>\alpha</math>a
|-
|5
|<math>\gamma</math>c
| <math>\beta</math>b →
|<math>\alpha</math>a
|-
|6
|<math>\gamma</math>c
|← b
|<math>\alpha</math>a <math>\beta</math>
|-
|7
|<math>\gamma</math>
|bc →
|<math>\alpha</math>a <math>\beta</math>
|-
|8
|<math>\gamma</math>
|← c
|<math>\alpha</math>a <math>\beta</math>b
|-©
|9
|
|<math>\gamma</math>c →
|<math>\alpha</math>a <math>\beta</math>b
|-
|(finish)
|
|
|<math>\alpha</math>a <math>\beta</math>b <math>\gamma</math>c
|}


==Variations==
This gives a differential relationship between the changes in temperature and volume, which can be integrated to find the invariant. The constant <math> k_B </math> is just a [[natural units| unit conversion factor]], which can be set equal to one:
An obvious generalization is to vary the number of jealous couples (or missionaries and cannibals), the capacity of the boat, or both.  If the boat holds 2 people, then 2 couples require 5 trips; with 4 or more couples, the problem has no solution.<ref>[http://www.sciencenews.org/view/generic/id/4512/title/Math_Trek__Tricky_Crossings Tricky Crossings], Ivars Peterson, ''Science News'', '''164''', #24 (December 13, 2003); accessed on line February 7, 2008, URL update March 12, 2011</ref>  If the boat can hold 3 people, then up to 5 couples can cross; if the boat can hold 4 people, any number of couples can cross.<ref name=b /><sup>,&nbsp;p.&nbsp;300.</sup>
:<math>\,
d(C_v N \log T) = - d( N \log V)
</math>


If an island is added in the middle of the river, then any number of couples can cross using a two-person boat.  If crossings from bank to bank are not allowed, then 8''n''&minus;6 one-way trips are required to ferry ''n'' couples across the river;<ref name=a /><sup>,&nbsp;p.&nbsp;76</sup> if they are allowed, then 4''n''+1 trips are required if ''n'' exceeds 4, although a minimal solution requires only 16 trips if ''n'' equals 4.<ref name=a /><sup>,&nbsp;p.&nbsp;79.</sup>  If the jealous couples are replaced by missionaries and cannibals, the number of trips required does not change if crossings from bank to bank are not allowed; if they are however the number of trips decreases to 4''n''&minus;1, assuming that ''n'' is at least 3.<ref name=a /><sup>,&nbsp;p.&nbsp;81.</sup>
So
:<math>\,
C_v N \log T + N \log V
</math>  


==History==
is an adiabatic invariant, which is related to the entropy
The first known appearance of the jealous husbands problem is in the medieval text ''[[Propositiones ad Acuendos Juvenes]]'', usually attributed to [[Alcuin]] (died 804.)  In Alcuin's formulation the couples are brothers and sisters, but the constraint is still the same&mdash;no woman can be in the company of another man unless her brother is present.<ref name=a /><sup>,&nbsp;p.&nbsp;74.</sup>  From the 13th to the 15th century, the problem became known throughout Northern Europe, with the couples now being husbands and wives.<ref name=b /><sup>,&nbsp;pp.&nbsp;291&ndash;293.</sup>  The problem was later put in the form of masters and valets; the formulation with missionaries and cannibals did not appear until the end of the 19th century.<ref name=a /><sup>,&nbsp;p.&nbsp;81</sup>  Varying the number of couples and the size of the boat was considered at the beginning of the 16th century.<ref name=b /><sup>,&nbsp;p.&nbsp;296.</sup>  Cadet de Fontenay considered placing an island in the middle of the river in 1879; this variant of the problem, with a two-person boat, was completely solved by Ian Pressman and [[David Singmaster]] in 1989.<ref name=a />
:<math>\,
S = C_v N \log T + N \log V - N \log N = N \log (T^{C_v} V/N)
</math>


==See also==
So entropy is an adiabatic invariant. The N log(N) term makes the entropy additive, so the entropy of two volumes of gas is the sum of the entropies of each one. 
*[[Fox, goose and bag of beans puzzle]]
 
*[[Transport puzzle]]
In a molecular interpretation, S is the logarithm of the phase space volume of all gas states with energy E(T) and volume V.
*[[Circumscription (logic)]]
 
For a monatomic ideal gas, this can easily be seen by writing down the energy,
 
:<math>E= {1\over 2m} \sum_k p_{k1}^2 + p_{k2}^2 + p_{k3}^2</math>
 
The different internal motions of the gas with total energy E define a sphere, the surface of a 3N-dimensional ball with radius <math>\scriptstyle \sqrt{2mE}</math>. The volume of the sphere is
 
:<math>{2\pi^{3N/2}(2mE)^{{3N-1}\over 2}}\over {\Gamma(3N/2)}</math>,
 
where <math>\Gamma</math> is the [[Gamma function]].
 
Since each gas molecule can be anywhere within the volume V, the volume in phase space occupied by the gas states with energy E is 
:<math>{2\pi^{3N/2}(2mE)^{{3N-1}\over 2}}V^N\over {\Gamma(3N/2)}</math>.
 
Since the N gas molecules are indistinguishable, the phase space volume is divided by <math>N! = \Gamma(N+1) </math>, the number of permutations of N molecules. 
 
Using [[Stirling's approximation]] for the gamma function, and ignoring factors that disappear in the logarithm after taking N large,
:<math>
S= N \big( 3/2 log(E)- 3/2 log(3N/2)+log(V)-log(N)\big ) </math>
:<math> = N \big( 3/2 log(\scriptstyle{\frac 2 3} \displaystyle E/N)+log(V/N)\big )</math>
 
Since the specific heat of a monatomic gas is 3/2, this is the same as the thermodynamic formula for the entropy.
 
=== Wien's Law — adiabatic expansion of a box of light ===
 
For a box of radiation, ignoring quantum mechanics, the energy of a classical field in thermal equilibrium is [[ultraviolet catastrophe|infinite]], since [[equipartition]] demands that each field mode has an equal energy on average and there are infinitely many modes. This is physically ridiculous, since it means that all energy leaks into high frequency electromagnetic waves over time.
 
Still, without quantum mechanics, there are some things that can be said about the equilibrium distribution from thermodynamics alone, because there is still a notion of adiabatic invariance that relates boxes of different size.
 
When a box is slowly expanded, the frequency of the light recoiling from the
wall can be computed from the [[Doppler shift]]. If the wall is not moving,
the light recoils at the same frequency. If the wall is moving slowly, the recoil frequency is only equal in the frame where the wall is stationary. In the frame where the wall is moving away from the light, the light coming in is bluer than the light coming out by twice the Doppler shift factor v/c.
:<math>
\Delta f = {2v\over c} f
</math>
 
On the other hand, the energy in the light is also decreased when the wall is moving away, because the light is doing work on the wall by radiation pressure. Because the light is reflected, the pressure is equal to twice the momentum carried by light, which is E/c. The rate at which the pressure does work on the wall is found by multiplying by the velocity:
:<math>\,
\Delta E = v{2E \over c}
</math>
 
This means that the change in frequency of the light is equal to the work done on the wall by the radiation pressure. The light that is reflected is changed both in frequency and in energy by the same amount:
:<math>
{\Delta f \over f} = {\Delta E \over E}
</math>
 
Since moving the wall slowly should keep a thermal distribution fixed, the probability that the light has energy E at frequency f must only be a function of E/f.
 
This function cannot be determined from thermodynamic reasoning alone, and Wien guessed at the form that was valid at high frequency. He supposed that the average energy in high frequency modes was suppressed by a Boltzmann-like factor. This is not the expected classical energy in the mode, which is <math>1/2\beta</math> by equipartition, but a new and unjustified assumption that fit the high-frequency data.
:<math>\,
<E_f> = e^{-\beta h f}
</math>
 
When the expectation value is added over all modes in a cavity, this is [[Wien approximation|Wien's distribution]], and it describes the thermodynamic distribution of energy in a classical gas of photons. Wien's Law implicitly assumes that light is statistically composed of packets that change energy and frequency in the same way. The entropy of a Wien gas scales as the volume to the power N, where N is the number of packets. This led Einstein to suggest that light is composed of localizable particles with energy proportional to the frequency. Then the entropy of the Wien gas can be given a statistical interpretation as the number of possible positions that the photons can be in.
 
== Classical mechanics — action variables ==
 
Suppose that a Hamiltonian is slowly time varying, for example, a one dimensional harmonic oscillator with a changing frequency.
:<math>
H_t(p,x) = {p^2\over 2m} + {m \omega(t)^2 x^2\over 2}
\,</math>
 
The [[action-angle variables|action]] J of a classical orbit is the area
enclosed by the orbit in phase space.
:<math>
J = \int_0^T p(t) {dx \over dt} dt
\,</math>
 
Since J is an integral over a full period, it is only a function of the energy. When
the Hamiltonian is constant in time and J is constant in time, the canonically conjugate variable <math>\theta</math> increases in time at a steady rate.
:<math>
{d\theta \over dt} = {\partial H \over \partial J} =H'(J)
\,</math>
 
So the constant H' can be used to change time derivatives along the orbit to partial derivatives with respect to <math>\theta</math> at constant J. Differentiating the integral for J with respect to J gives an identity that fixes H':
:<math>
{dJ\over dJ } = 1 = \int_0^T \bigg( {\partial p \over \partial J} {dx \over dt}
+ p {\partial \over \partial J} {dx \over dt} \bigg) dt =
H' \int_0^T \bigg({\partial p \over \partial J}{\partial x \over \partial \theta} - {\partial p \over \partial \theta}{\partial x \over \partial J}\bigg) dt
\,</math>
 
The integrand is the [[Poisson bracket]] of x and p. The Poisson bracket of two canonically conjugate quantities like x and p is equal to 1 in any canonical coordinate system. So
 
<math>
1 = H' \int_0^T \{ x,p \} dt = H' T
\,</math>
 
and H' is the inverse period. The variable <math>\theta</math> increases by an equal amount in each period for all values of J— it is an angle-variable.
 
; Adiabatic Invariance of J
 
The Hamiltonian is a function of J only, and in the simple case of the Harmonic oscillator.
:<math>\,
H= \omega J
\,</math>
 
When H has no time dependence, J is constant. When H is slowly time varying, the
rate of change of J can be computed by reexpressing the integral for J
:<math>
J = \int_0^{2\pi} p {\partial x \over \partial \theta} d\theta
\,</math>
 
The time derivative of this quantity is
:<math>
{dJ\over dt} = \int_0^{2\pi} \bigg({dp \over dt} {\partial x\over \partial \theta} +
p {d\over dt} {\partial x \over \partial \theta} \bigg) d\theta
\,</math>
 
Replacing time derivatives with theta derivatives,
:<math>
{dJ \over dt} = \int_0^{2\pi} \bigg({\partial p \over \partial \theta} {\partial x \over \partial \theta} + p {\partial \over \partial \theta} {\partial x \over \partial \theta} \bigg) d\theta
\,</math>
 
So as long as the coordinates J,<math>\theta</math> do not change appreciably over one period, this expression can be integrated by parts to give zero. This means
that for slow variations, there is no lowest order change in the area enclosed by
the orbit. This is the adiabatic invariance theorem— the action variables are adiabatic invariants.
 
For a harmonic oscillator, the area in phase space of an orbit at energy E is the area
of the ellipse of constant energy,
:<math>
E = {p^2\over 2m} + {m\omega^2 x^2\over 2}
\,</math>
 
The x-radius of this ellipse is <math>\scriptstyle \sqrt{2E/\omega^2m}</math>, while the p-radius of the ellipse is <math>\scriptstyle \sqrt{2mE}</math>. Multiplying, the area is <math>2\pi E/\omega</math>. So if a pendulum is slowly drawn in, so that the frequency changes, the energy changes by a proportional amount.
 
=== Old quantum theory ===
 
After Planck identified that Wien's law can be extended to all frequencies, even very low ones, by interpolating with the classical equipartition law for radiation, physicists wanted to understand the quantum behavior of other systems.
 
The Planck radiation law quantized the motion of the field oscillators in units of energy proportional to the frequency:
:<math>
E= h f = \hbar \omega
\,</math>
 
This is the only sensible quantization. The quantum can only depend on the energy/frequency by adiabatic invariance, and since the energy must be additive when putting boxes end to end, the levels must be equally spaced.
 
Einstein, followed by Debye, extended the domain of quantum mechanics by considering the sound modes in a solid as [[Einstein solid|quantized oscillators]]. This model explained why the specific heat of solids approached zero at low temperatures,
instead of staying fixed at <math>3k_B</math> as predicted by classical [[equipartition theorem| equipartition]].
 
At the [[Solvay conference]], the question of quantizing other motions was raised, and [[Hendrik Lorentz|Lorentz]] pointed out a problem. If you consider a quantum pendulum whose string is shortened very slowly, the quantum number of the pendulum cannot change because at no point is there a high enough frequency to cause a transition between the states. But the frequency of the pendulum changes when the string is shorter, so the quantum states change energy.
 
Einstein responded that for slow pulling, the frequency and energy of the pendulum both change but the ratio stays fixed. This is analogous to Wien's observation that under slow motion of the wall the energy to frequency ratio of reflected waves is constant. The conclusion was that the quantities to quantize must be adiabatic invariants.
 
This line of argument was extended by Sommerfeld into a general theory: the quantum number of an arbitrary mechanical system is given by the adiabatic action variable. Since the action variable in the harmonic oscillator is an integer, the general condition is:
:<math>
\int p dq = n h
\,</math>
 
This condition was the foundation of the [[old quantum theory]], which was able to predict the qualitative behavior of atomic systems. The theory is inexact for small quantum numbers, since it mixes classical and quantum concepts. But it was a useful half-way step to the [[matrix mechanics|new quantum theory]].
 
== Plasma physics ==
 
In [[plasma physics]] there are three adiabatic invariants of charged particle motion.
 
=== The first adiabatic invariant, &mu; ===
The '''magnetic moment''' of a gyrating particle,
:<math>\mu = \frac{\frac{1}{2}mv_\perp^2}{B}</math>,
 
is a constant of the motion to all orders in an expansion in <math>\omega/\omega_c</math>, where <math>\omega</math> is the rate of any changes experienced by the particle, e.g., due to collisions or due to temporal or spatial variations in the magnetic field. Consequently the magnetic moment remains nearly constant even for changes at rates approaching the gyrofrequency. When μ is constant, the perpendicular particle energy is proportional to ''B'', so the particles can be heated by increasing ''B'', but this is a 'one shot' deal because the field cannot be increased indefinitely.
 
There are some important situations in which the magnetic moment is ''not'' invariant:
* '''Magnetic pumping:''' If the collision frequency is larger than the pump frequency, μ is no longer conserved. In particular, collisions allow net heating by transferring some of the perpendicular energy to parallel energy.
* '''Cyclotron heating:''' If ''B'' is oscillated at the cyclotron frequency, the condition for adiabatic invariance is violated and heating is possible. In particular, the induced electric field rotates in phase with some of the particles and continuously accelerates them.
* '''Magnetic cusps:''' The magnetic field at the center of a cusp vanishes, so the cyclotron frequency is automatically smaller than the rate of ''any'' changes. Thus the magnetic moment is not conserved and particles are scattered relatively easily into the [[Magnetic mirror|loss cone]].
 
=== The second adiabatic invariant, ''J'' ===
The '''longitudinal invariant''' of a particle trapped in a magnetic mirror,
:<math>J = \int_a^b v_\|\, ds</math>,
 
where the integral is between the two turning points, is also an adiabatic invariant. This guarantees, for example, that a particle in the [[magnetosphere]] moving around the Earth always returns to the same line of force. The adiabatic condition is violated in '''transit-time magnetic pumping''', where the length of a magnetic mirror is oscillated at the bounce frequency, resulting in net heating.
 
=== The third adiabatic invariant, &Phi; ===
The total magnetic flux Φ enclosed by a drift surface is the third adiabatic invariant, associated with the periodic motion of mirror-trapped particles drifting around the axis of the system. Because this drift motion is relatively slow, Φ is often not conserved in practical applications.


==References==
==References==
{{reflist}}
* {{cite book|last=Yourgrau|first=Wolfgang|coauthors=Stanley Mandelstam|title=Variational Principles in Dynamics and Quantum Theory|publisher=Dover|location=New York|year=1979|isbn=0-486-63773-5}} §10
* {{cite book|last=Pauli|first=Wolfgang|editor=Charles P. Enz|title=Pauli Lectures on Physics Vol. 4|publisher=MIT Press|location=Cambridge, Mass|year=1973|isbn=0-262-66035-0}} pp. 85-89
 
==External links==
* [http://farside.ph.utexas.edu/teaching/plasma/lectures/node25.html lecture notes on the second adiabatic invariant]
* [http://farside.ph.utexas.edu/teaching/plasma/lectures/node26.html lecture notes on the third adiabatic invariant]


[[Category:Logic puzzles]]
[[Category:Classical mechanics]]
[[Category:Concepts in physics]]
[[Category:Quantum mechanics]]
[[Category:Thermodynamics]]
[[Category:Plasma physics]]


[[pt:Problema dos canibais e missionários]]
[[lt:Adiabatinis invariantas]]
[[uk:Задача про місіонерів та канібалів]]
[[ja:断熱不変量]]
[[zh:传教士和吃人恶魔问题]]
[[ru:Адиабатический инвариант]]
[[sk:Adiabatický invariant]]
[[uk:Адіабатичний інваріант]]
[[zh:绝热不变量]]

Revision as of 17:22, 13 August 2014

An adiabatic invariant is a property of a physical system that stays constant when changes occur slowly.

In thermodynamics, an adiabatic process is a change that occurs without heat flow, and slowly compared to the time to reach equilibrium. In an adiabatic process, the system is in equilibrium at all stages. Under these conditions, the entropy is constant.

In mechanics, an adiabatic change is a slow deformation of the Hamiltonian, where the fractional rate of change of the energy is much slower than the orbital frequency. The area enclosed by the different motions in phase space are the adiabatic invariants.

In quantum mechanics, an adiabatic change is one that occurs at a rate much slower than the difference in frequency between energy eigenstates. In this case, the energy states of the system do not make transitions, so that the quantum number is an adiabatic invariant.

The old quantum theory was formulated by equating the quantum number of a system with its classical adiabatic invariant. This determined the form of the Bohr-Sommerfeld quantization rule: the quantum number is the area in phase space of the classical orbit.

Thermodynamics

In thermodynamics, adiabatic changes are those that do not increase the entropy. They occur slowly, and allow heat flow only between objects at the same temperature. For isolated systems, an adiabatic change allows no heat to flow in or out.

Adiabatic expansion of an ideal gas

If a container with an ideal gas is expanded instantaneously, the temperature of the gas doesn't change at all, because none of the molecules slow down. The molecules keep their kinetic energy, but now the gas occupies a bigger volume. If the container expands slowly, however, so that the ideal gas pressure law holds at any time, gas molecules lose energy at the rate that they do work on the expanding wall. The amount of work they do is the pressure times the area of the wall times the outward displacement, which is the pressure times the change in the volume of the gas:

If no heat enters the gas, the energy in the gas molecules is decreasing by the same amount. By definition, a gas is ideal when its temperature is only a function of the internal energy per particle, not the volume. So

Where is the specific heat at constant volume. When the change in energy is entirely due to work done on the wall, the change in temperature is given by:

This gives a differential relationship between the changes in temperature and volume, which can be integrated to find the invariant. The constant is just a unit conversion factor, which can be set equal to one:

So

is an adiabatic invariant, which is related to the entropy

So entropy is an adiabatic invariant. The N log(N) term makes the entropy additive, so the entropy of two volumes of gas is the sum of the entropies of each one.

In a molecular interpretation, S is the logarithm of the phase space volume of all gas states with energy E(T) and volume V.

For a monatomic ideal gas, this can easily be seen by writing down the energy,

The different internal motions of the gas with total energy E define a sphere, the surface of a 3N-dimensional ball with radius . The volume of the sphere is

,

where is the Gamma function.

Since each gas molecule can be anywhere within the volume V, the volume in phase space occupied by the gas states with energy E is

.

Since the N gas molecules are indistinguishable, the phase space volume is divided by , the number of permutations of N molecules.

Using Stirling's approximation for the gamma function, and ignoring factors that disappear in the logarithm after taking N large,

Since the specific heat of a monatomic gas is 3/2, this is the same as the thermodynamic formula for the entropy.

Wien's Law — adiabatic expansion of a box of light

For a box of radiation, ignoring quantum mechanics, the energy of a classical field in thermal equilibrium is infinite, since equipartition demands that each field mode has an equal energy on average and there are infinitely many modes. This is physically ridiculous, since it means that all energy leaks into high frequency electromagnetic waves over time.

Still, without quantum mechanics, there are some things that can be said about the equilibrium distribution from thermodynamics alone, because there is still a notion of adiabatic invariance that relates boxes of different size.

When a box is slowly expanded, the frequency of the light recoiling from the wall can be computed from the Doppler shift. If the wall is not moving, the light recoils at the same frequency. If the wall is moving slowly, the recoil frequency is only equal in the frame where the wall is stationary. In the frame where the wall is moving away from the light, the light coming in is bluer than the light coming out by twice the Doppler shift factor v/c.

On the other hand, the energy in the light is also decreased when the wall is moving away, because the light is doing work on the wall by radiation pressure. Because the light is reflected, the pressure is equal to twice the momentum carried by light, which is E/c. The rate at which the pressure does work on the wall is found by multiplying by the velocity:

This means that the change in frequency of the light is equal to the work done on the wall by the radiation pressure. The light that is reflected is changed both in frequency and in energy by the same amount:

Since moving the wall slowly should keep a thermal distribution fixed, the probability that the light has energy E at frequency f must only be a function of E/f.

This function cannot be determined from thermodynamic reasoning alone, and Wien guessed at the form that was valid at high frequency. He supposed that the average energy in high frequency modes was suppressed by a Boltzmann-like factor. This is not the expected classical energy in the mode, which is by equipartition, but a new and unjustified assumption that fit the high-frequency data.

When the expectation value is added over all modes in a cavity, this is Wien's distribution, and it describes the thermodynamic distribution of energy in a classical gas of photons. Wien's Law implicitly assumes that light is statistically composed of packets that change energy and frequency in the same way. The entropy of a Wien gas scales as the volume to the power N, where N is the number of packets. This led Einstein to suggest that light is composed of localizable particles with energy proportional to the frequency. Then the entropy of the Wien gas can be given a statistical interpretation as the number of possible positions that the photons can be in.

Classical mechanics — action variables

Suppose that a Hamiltonian is slowly time varying, for example, a one dimensional harmonic oscillator with a changing frequency.

The action J of a classical orbit is the area enclosed by the orbit in phase space.

Since J is an integral over a full period, it is only a function of the energy. When the Hamiltonian is constant in time and J is constant in time, the canonically conjugate variable increases in time at a steady rate.

So the constant H' can be used to change time derivatives along the orbit to partial derivatives with respect to at constant J. Differentiating the integral for J with respect to J gives an identity that fixes H':

The integrand is the Poisson bracket of x and p. The Poisson bracket of two canonically conjugate quantities like x and p is equal to 1 in any canonical coordinate system. So

and H' is the inverse period. The variable increases by an equal amount in each period for all values of J— it is an angle-variable.

Adiabatic Invariance of J

The Hamiltonian is a function of J only, and in the simple case of the Harmonic oscillator.

When H has no time dependence, J is constant. When H is slowly time varying, the rate of change of J can be computed by reexpressing the integral for J

The time derivative of this quantity is

Replacing time derivatives with theta derivatives,

So as long as the coordinates J, do not change appreciably over one period, this expression can be integrated by parts to give zero. This means that for slow variations, there is no lowest order change in the area enclosed by the orbit. This is the adiabatic invariance theorem— the action variables are adiabatic invariants.

For a harmonic oscillator, the area in phase space of an orbit at energy E is the area of the ellipse of constant energy,

The x-radius of this ellipse is , while the p-radius of the ellipse is . Multiplying, the area is . So if a pendulum is slowly drawn in, so that the frequency changes, the energy changes by a proportional amount.

Old quantum theory

After Planck identified that Wien's law can be extended to all frequencies, even very low ones, by interpolating with the classical equipartition law for radiation, physicists wanted to understand the quantum behavior of other systems.

The Planck radiation law quantized the motion of the field oscillators in units of energy proportional to the frequency:

This is the only sensible quantization. The quantum can only depend on the energy/frequency by adiabatic invariance, and since the energy must be additive when putting boxes end to end, the levels must be equally spaced.

Einstein, followed by Debye, extended the domain of quantum mechanics by considering the sound modes in a solid as quantized oscillators. This model explained why the specific heat of solids approached zero at low temperatures, instead of staying fixed at as predicted by classical equipartition.

At the Solvay conference, the question of quantizing other motions was raised, and Lorentz pointed out a problem. If you consider a quantum pendulum whose string is shortened very slowly, the quantum number of the pendulum cannot change because at no point is there a high enough frequency to cause a transition between the states. But the frequency of the pendulum changes when the string is shorter, so the quantum states change energy.

Einstein responded that for slow pulling, the frequency and energy of the pendulum both change but the ratio stays fixed. This is analogous to Wien's observation that under slow motion of the wall the energy to frequency ratio of reflected waves is constant. The conclusion was that the quantities to quantize must be adiabatic invariants.

This line of argument was extended by Sommerfeld into a general theory: the quantum number of an arbitrary mechanical system is given by the adiabatic action variable. Since the action variable in the harmonic oscillator is an integer, the general condition is:

This condition was the foundation of the old quantum theory, which was able to predict the qualitative behavior of atomic systems. The theory is inexact for small quantum numbers, since it mixes classical and quantum concepts. But it was a useful half-way step to the new quantum theory.

Plasma physics

In plasma physics there are three adiabatic invariants of charged particle motion.

The first adiabatic invariant, μ

The magnetic moment of a gyrating particle,

,

is a constant of the motion to all orders in an expansion in , where is the rate of any changes experienced by the particle, e.g., due to collisions or due to temporal or spatial variations in the magnetic field. Consequently the magnetic moment remains nearly constant even for changes at rates approaching the gyrofrequency. When μ is constant, the perpendicular particle energy is proportional to B, so the particles can be heated by increasing B, but this is a 'one shot' deal because the field cannot be increased indefinitely.

There are some important situations in which the magnetic moment is not invariant:

  • Magnetic pumping: If the collision frequency is larger than the pump frequency, μ is no longer conserved. In particular, collisions allow net heating by transferring some of the perpendicular energy to parallel energy.
  • Cyclotron heating: If B is oscillated at the cyclotron frequency, the condition for adiabatic invariance is violated and heating is possible. In particular, the induced electric field rotates in phase with some of the particles and continuously accelerates them.
  • Magnetic cusps: The magnetic field at the center of a cusp vanishes, so the cyclotron frequency is automatically smaller than the rate of any changes. Thus the magnetic moment is not conserved and particles are scattered relatively easily into the loss cone.

The second adiabatic invariant, J

The longitudinal invariant of a particle trapped in a magnetic mirror,

,

where the integral is between the two turning points, is also an adiabatic invariant. This guarantees, for example, that a particle in the magnetosphere moving around the Earth always returns to the same line of force. The adiabatic condition is violated in transit-time magnetic pumping, where the length of a magnetic mirror is oscillated at the bounce frequency, resulting in net heating.

The third adiabatic invariant, Φ

The total magnetic flux Φ enclosed by a drift surface is the third adiabatic invariant, associated with the periodic motion of mirror-trapped particles drifting around the axis of the system. Because this drift motion is relatively slow, Φ is often not conserved in practical applications.

References

  • 20 year-old Real Estate Agent Rusty from Saint-Paul, has hobbies and interests which includes monopoly, property developers in singapore and poker. Will soon undertake a contiki trip that may include going to the Lower Valley of the Omo.

    My blog: http://www.primaboinca.com/view_profile.php?userid=5889534 §10
  • 20 year-old Real Estate Agent Rusty from Saint-Paul, has hobbies and interests which includes monopoly, property developers in singapore and poker. Will soon undertake a contiki trip that may include going to the Lower Valley of the Omo.

    My blog: http://www.primaboinca.com/view_profile.php?userid=5889534 pp. 85-89

External links

lt:Adiabatinis invariantas ja:断熱不変量 ru:Адиабатический инвариант sk:Adiabatický invariant uk:Адіабатичний інваріант zh:绝热不变量