The Variational Principle

3.1: Fermat’s Principle

Fermat’s principles of least time: light travels between two points alone the path that requires the least time, as compared to other nearby paths.

Snell’s Law:

n1sin(θ1)=n2sin(θ2)\begin{equation} n_1\sin(\theta_1)=n_2\sin(\theta_2) \end{equation}

The refraction index of a medium nn is defined as n(r)=C/vn(r)=C/v, where CC is the speed of light in vacuum and vv is the speed of light in the new medium. Then, speed of light in a medium is equal to v=C/n(r)v=C/n(r), where n(x,r,z)=n(r)n(x,r,z)=n(r) can be a continuous function. The time required to travel a distance dsds at speed vv is given by:

dt=distancespeed=dsvdt=\frac{\text{distance}}{\text{speed}}=\frac{ds}{v}

The total time required to travel from aa to bb is given by:

t=dt=tdC/n(r)=1Cdsn(r)\begin{equation} t=\int dt=t\int\frac{d}{C/n(r)}=\frac{1}{C}\int \frac{ds}{n(r)} \end{equation}

💡
Fermat’s principles of least time: light travels between two points alone the path that requires the least time, as compared to other nearby paths.

One question raises from Fermat’s principle of least time is: how do light rays know which path is closest ? The equation is answered by a method called the variation of calculus.

3.2: The Calculus of Variation

From the above section, we know that the time required to travel from points AA to BB is given by:

t=1Cn(r)dst=\frac{1}{C}\int n(r) {ds}

In the integral, dsds can be expressed in terms of dxdx and dydy, giving us ds=dx2+dy2ds=\sqrt{dx^2+dy^2}. The refraction index n(r)=n(x)n(r)=n(x) because we are integrating alone that xx coordinate. Then, the integral becomes:

t=1Cn(x)ds=1Cn(x)dx2+dy2=1Cn(x)1+dy2dx2dx=1Cn(x)1+y2dxt=\frac{1}{C}\int n(x) {ds}=\frac{1}{C}\int n(x)\sqrt{dx^2+dy^2}=\frac{1}{C}\int n(x)\sqrt{1+\frac{dy^2}{dx^2}}dx=\frac{1}{C}\int n(x)\sqrt{1+y'^2} dx
The picture explains the integral D(f)D(f), the distance from AA to BB alone ff, where xx coordinates varies from x1x_1 to x2x_2. Because we are integrating alone xx coordinates (we use xx coordinates to define/constrain the path), n(r)=n(x)n(r)=n(x)

In this case, the integrand depends both on yy’ and dxdx, so the integral is rewritten as:

t=1Cn(x,y)1+y2dx=F(x,y(x),y(x))dxt=\frac{1}{C}\int n(x,y)\sqrt{1+y'^2} dx=\int F(x, y(x), y'(x))dx

where

1Cn(x,y)1+y2=F(x,y(x),y(x))\frac{1}{C}n(x,y)\sqrt{1+y'^2} =F(x, y(x), y'(x))

Example: Light Path Between Two Points in Glass

Consider first a very simple special case. Suppose the index of refraction n is constant throughout a sheet of glass, and that the endpoint of a light ray at x=x0x = x_0, y=y0y = y_0 is directly across the sheet from the beginning point at x=0x = 0, y=y0y = y_0. Then the time to penetrate the sheet is what?


3.21: Euler’s Method Derivation Part 1

More generally, Euler and Lagrange considered some arbitrary integrals in the form:

I=F(x,y(x),y(x))I=F(x, y(x), y'(x))

and wanted to solve paths y(x)y(x) that maximum and minimized. In other words, they wanted to find the stationary points where I=0I'=0.

Suppose we have a function f(x1,x2,)=f(xi)f(x_1, x_2, \cdots)=f(x_i) of independent variables i=1,2,3,,Ni=1,2,3,\cdots, N, and we are asked to find the stationary points. At stationary points, f(xi)=0f’(x_i)=0, so f(xi)f(x_i)  must be a constant. (This might sounds confusing at the first point. Consider functions f(xi)=3xandf(xi)=3f(x_i)=3x\:\text{and}\: f(x_i)=3, their derivatives f(xi)=30andf(xi)=0f’(x_i)=3\ne 0\:\text{and}\:f’(x_i)=0 f(xi)f(x_i) must be a constant. Thus, f(xi)f(x_i) must be a constant for f(xi)=0f’(x_i)=0). We represent a horizontal shift in f(xi)f(x_i) as f(xi+δxi)f(x_i+\delta x_i). When reaching a stationary point, the function is a constant, so f(xi)f(xi+δxi)f(x_i)\approx f(x_i+\delta x_i), where δxi0\delta x_i\rightarrow 0. Using Taylor expansion, f(xi+δxi)f(x_i+\delta x_i) is equal to:

f(xi+δxi)=f(xi)+fxjδxj+12!f2xjxkδxjδxk+f(x_i+\delta x_i)=f(x_i)+\frac{\partial f}{\partial x_j}\delta x_j+\frac{1}{2!}\frac{\partial f}{\partial^2 x_jx_k}\delta x_j\delta x_k+\cdots

In order to the constant for f(xi+δxi)=f(xi)f(x_i+\delta x_i)=f(x_i), all other terms need to be 0. This is only possible if f/xj=0\partial f/\partial x_j=0. Thus, when y(x)y(x) reaches a stationary point,

fxj=0withj=1,2,,N\begin{equation} \boxed{\frac{\partial f}{\partial x_j}=0\quad\text{with}\quad j=1,2,\cdots, N} \end{equation}

3.22: Euler’s Equation Derivation Part 2

In order to find such paths, we introduce some new functions ϵ\epsilon and η\eta. Suppose yˉ(x)\bar{y}(x) is a path for stationary points and y(x){y}(x) is a path nearby. The path yˉ(x)\bar{y}(x) can be expressed using y(x)y(x) plus a small difference.

y(x)=yˉ(x)+ϵη{y}(x)=\bar{y}(x)+\epsilon\eta

At stationary points, I=0I’=0. Applying chain rule for multivariable functions, we integrate II with respect to ϵ\epsilon, giving us:

0=dIdϵ=ddϵ(F(x,y(x),y(x))dx=Fxxϵ+Fy(x)y(x)ϵ+Fy(x)y(x)ϵdx\begin{align*} 0&=\frac{dI}{d\epsilon}=\int\frac{d}{d\epsilon}(F(x, y(x), y'(x)) dx\\ &=\int \frac{\partial F}{x}\frac{\partial x}{\partial\epsilon}+\frac{\partial F}{y(x)}\frac{\partial y(x)}{\partial\epsilon}+\frac{\partial F}{\partial y'(x)}\frac{\partial y'(x)}{\partial\epsilon} dx \end{align*}

Because f/xj=0\partial f/\partial x_j=0, the expression becomes:

0=Fy(x)y(x)ϵ+Fy(x)y(x)ϵ0=\int \frac{\partial F}{y(x)}\frac{\partial y(x)}{\partial\epsilon}+\frac{\partial F}{\partial y'(x)}\frac{\partial y'(x)}{\partial\epsilon}

Integrating y(x)y(x) and y(x)y’(x) with respect to ϵ\epsilon, we get:

y(x)ϵ=ηy(x)ϵ=η\frac{\partial y(x)}{\epsilon}=\eta\quad\frac{\partial y'(x)}{\partial\epsilon}=\eta'

Then, the integral becomes:

0=Fy(x)η+Fy(x)ηdx0=\int \frac{\partial F}{\partial y(x)}\eta +\frac{\partial F}{\partial y'(x)}\eta'dx

Using integrating by parts, we get:

Fy(x)ηdx=Fy(x)ηdxηddx(Fy(x))dx=Fy(x)ηηddx(Fy(x))dx\int \frac{\partial F}{\partial y'(x)}\eta' dx=\frac{\partial F}{\partial y'(x)}\int \eta' dx-\int \eta\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right) dx=\frac{\partial F}{\partial y'(x)}\eta-\int \eta\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right) dx

Because η(x)=0\eta(x)=0, the equation is equal to:

Fy(x)ηdx=ηddx(Fy(x))dx\int \frac{\partial F}{\partial y'(x)}\eta' dx=-\int \eta\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right) dx

Then, the integral becomes:

0=Fy(x)ηdxηddx(Fy(x))dx=(Fy(x)dxddx(Fy(x)))ηdx0=\int \frac{\partial F}{\partial y(x)}\eta dx-\int \eta\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right) dx=\int \left (\frac{\partial F}{\partial y(x)}dx-\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right)\right) \eta dx

giving us

Fy(x)=ddx(Fy(x))\begin{equation} \boxed{\frac{\partial F}{\partial y(x)}=\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right)} \end{equation}

The integrand is known us the Euler’s equation.

0=Fy(x)ddx(Fy(x))\begin{equation} \boxed{0=\frac{\partial F}{\partial y(x)}-\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right)} \end{equation}

💡
This might be nuance, but we are getting used to y=dydty’=dy’dt. Note that in Euler’s equation, y=dy/dxy’=dy/dx.

Example: The Straight Wire

Minimize the equation

s=ds=dx2+dy2=1+y2dxs=\int{ds}=\int \sqrt{dx^2+dy^2}=\int \sqrt{1+y'^2} dx

3.3: Geodesics

One application of the calculus of variation is to find geodesics, which are the stationary (usually shortest) paths between two points on a given surface. Before diving into details, we will talk about cylindrical and spherical coordinates in Cartesian and polar forms.

3.31: Coordinates

Cylindrical Coordinates

Cartesin(rcosθ,rsinθ,z)\text{Cartesin}\quad (r\cos\theta, r\sin\theta, z)

In polar form

Polars=(rθ,z)\text{Polar}\quad s=(r\theta,z)
Polards=(rdθ,z)\text{Polar}\quad d\vec{s}=(rd\theta,z)
Polards=r2dθ2+z2\text{Polar}\quad d{s}=\sqrt{r^2d\theta^2+z^2}

Spherical coordinates

By pythagorean theorem, the length of the red line is given by:

s=x2+y2+z2s=\sqrt{x^2+y^2+z^2}

where:

x=rsinθcosϕy=rsinθsinϕz=rcosϕx=r\sin\theta\cos\phi \quad y=r\sin\theta\sin\phi \quad z=r\cos\phi


For spherical coordinates, we divide a sphere into infinitesimal number of cubes.

s=rdθs=rd\theta

Example: Geodesics on a Sphere

Consider the problem of finding the shortest distance between two points on the surface of a sphere. The sphere has a radius of RR. Find the geodesics on a sphere


3.4: Brachistochrone

3.41: A Cycloid

The section is not included in the textbook, but it will be beneficial to know some math on cycloids for our understanding of the brachistochrone problem.

Cycloids

Cups:

A graph of a cycloid.

In polar forms, a cycloid has equations of:

x=r(θsinθ)y=r(1cosθ)x=r(\theta-\sin\theta)\quad y=r(1-\cos\theta)

3.4.2: Time and Paths Taken for a Brachistochrone

The problem is to find the shape of a track between two given points, such that a small ball starting at rest at the upper point-and sliding without friction alone the track under the influence of gravity-arrives at the lower point in the shortest time.

Brachistochrone: The path between two points AA and BB which minimizes the time taking by a particle falling from AA to BB under the influence of gravity.

t=dsvt=\int \frac{ds}{v}

Let point AA be the origin of the coordinate system we define. Because the particle is dropped, it has an initial velocity and kinetic energy of zero. As the point is zero, gravitational potential energy is also zero. Thus, the initial, which is equal to the total, energy is zero. By energy conservation, we get:

E=12mv2mgy=0E=\frac{1}{2}mv^2-mgy=0

The distance dsds is given by:

ds=dx2+dy2=1+y2dxds=\sqrt{dx^2+dy^2}=\sqrt{1+y'^2}dx

Then, time tt is equal to:

t=1+y22gydxort=1+x22gydy\begin{equation} t=\int \frac{\sqrt{1+y'^2}}{\sqrt{2gy}}dx\quad \text{or}\quad t=\int \frac{\sqrt{1+x'^2}}{\sqrt{2gy}}dy \end{equation}

Using Euler’s equation, t=t(y,x,x)t=t(y, x', x) and F()=(1+x2)/(2gy)F()=(1+x'^2)/(2gy). Then,

0=FxddyFx0=\frac{\partial F}{\partial x}-\frac{d}{dy}\frac{\partial F}{\partial x'}

Because FF is independent of xx, F/x=0\partial F/\partial x=0, giving us:

Fx=12gyx1+x2=k\frac{\partial F}{\partial x'}=\frac{1}{\sqrt{2gy}}\frac{x'}{\sqrt{1+x'^2}}=k

Solving for xx', we get:

x=±2gk2y12k2gy=±y1/2gk2yyayx'=\pm\sqrt{\frac{2gk^2y}{1-2k^2gy}}=\pm\sqrt{\frac{y}{1/2gk^2-y}}\equiv\sqrt{\frac{y}{a-y}}

We choose the plus square root. Integrating over yy, we get:

x=dx=yaydyx=\int dx=\int \sqrt{\frac{y}{a-y}}dy

Note that we integrate with respect to yy because x=x(y)x'=x'(y) is a function of yy, where yy is the dependent variable. The integral is evaluated using substitution. Let y=asin2(θ/2)=a/2(1cosθ)y=a\sin^2(\theta/2)=a/2(1-\cos\theta), the integral is equal to:

x=a2(θsinθ)x=\frac{a}{2}(\theta-\sin\theta)

Differentiating xx with respect to θ\theta, we get

x=a2(1cosθ)x'=\frac{a}{2}(1-\cos\theta)

Expressing dydy in term of θ\theta, we get:

dydθ=a2(sinθ)\frac{dy}{d\theta}=\frac{a}{2}(\sin\theta)
dy=a2(sinθ)dθdy=\frac{a}{2}(\sin\theta)d\theta

Using separable integration, yy is equal to:

y=a2(1cosθ)y=\frac{a}{2}(1-\cos\theta)

Notice that both xx and yy represent equations for a cycloid, the curve traced by a point on a circle as it rolls alone a straight line without slipping. The quantity aa and final angle θb\theta_b can be determined from coordinates (xb,yb)(x_b, y_b) of the final position.

A graph of a cycloid. Note that top points of a cycloid represents a half of a complete revolution, where θ=π\theta=\pi. In order to see this, imaging a point on ground initially, when the points is at top, the circle has completed one half of a revolution, so θ=π\theta=\pi.

Substitute xx or yy into (6), the time tt required to fall to the final position is:

t=1+x22gydy=12g1+a2(1cosθ)2/4a(1cosθ)/2a2(1+sinθ)dθt=\int \sqrt{\frac{1+x'^2}{2gy}}dy=\frac{1}{\sqrt{2g}}\int \sqrt{\frac{1+a^2(1-\cos\theta)^2/4}{a(1-\cos\theta)/2}}\frac{a}{2}(1+\sin\theta)d\theta
t=a2g0θfdθ=a2gθf\begin{equation} \boxed{t=\sqrt{\frac{a}{2g}}\int^{\theta_f}_0d\theta=\sqrt{\frac{a}{2g}}\theta_f} \end{equation}

Notice that if (xb,yb)=(aπ/2,a)(x_b, y_b)=(a\pi/2, a), time tt is equal to:

t=πa2gt=\pi\sqrt{\frac{a}{2g}}

Example: Path Traveled by Lights in Atmosphere

We return to where we began the chapter, with Fermat’s principle of stationary time, illustrated in Figure 3.9(a). Bringing to bear the calculus of variations, we can now find the path of a light ray in a medium like earth’s atmosphere, where the index of refraction n is a continuous function of position. If a ray of light from a star descends through the atmosphere, it encounters an increasing density and an increasing index of refraction. We might therefore expect the ray to bend continuously, entering the atmosphere at some angle θaθ_a and reaching the ground at a steeper angle θbθ_b . For simplicity, take the earth to be essentially flat over the horizontal range of the ray and assume the index of refraction n=n(y)n = n(y) only, where yy is the vertical direction. The light travel time is then equal to what?


3.5 Several Dependent Variables

Fyi(x)ddx(Fyi(x))=0\boxed{\frac{\partial F}{\partial y_i(x)}-\frac{d}{dx}\left(\frac{\partial F}{\partial y_i'(x)}\right)=0}

Example: Geodesics in Three Dimension

Using Euler’s equation for several dependent variable, find geodesics in three dimensions.


3.6: Mechanics From a Variation Principle

Recall from relativity that proper time (the reading in the frame moving with the clock) is equal to τ=Δtγ\tau=\Delta t\gamma, where Δt\Delta t is the reading in another frame (For instance, earth’s frame).

Δt=τγ\Delta t=\frac{\tau}{\gamma}

We want to find a general formulation of mechanics that is based entirely on a variational principle. Consider a relativistic particle, the proper time in the particle’s frame is τ\tau, and the reading in earth’s frame is Δt\Delta t.

I=dτ=dtγI=\int d\tau=\int \frac{dt}{\gamma}

Using space-time four dimensional coordinates (c,x,y,z)(c,x,y,z), we know that

I=dτ=dtγ=dt1x˙cy˙cz˙cI=\int d\tau=\int \frac{dt}{\gamma}=\int dt\sqrt{1-\frac{\dot x}{c}-\frac{\dot y}{c}-\frac{\dot z}{c}}

Using variation of calculus for several calculus we derived in section 3.6, we get:

FxddtFx=0,FyddtFy=0,FzddtFz=0\frac{\partial F}{\partial x}-\frac{d}{dt}\frac{\partial F}{\partial x'}=0, \quad \frac{\partial F}{\partial y}-\frac{d}{dt}\frac{\partial F}{\partial y'}=0, \quad \frac{\partial F}{\partial z}-\frac{d}{dt}\frac{\partial F}{\partial z'}=0

Because the FF is independent of x,yx, y and zz, expression becomes:

ddtFx=0,ddtFy=0,ddtFz=0\frac{d}{dt}\frac{\partial F}{\partial x'}=0, \quad \frac{d}{dt}\frac{\partial F}{\partial y'}=0, \quad \frac{d}{dt}\frac{\partial F}{\partial z'}=0
ddt(γx˙)=0,ddt(γy˙)=0,ddt(γz˙)=0\frac{d}{dt}(\gamma\dot{x})=0, \quad \frac{d}{dt}(\gamma\dot{y})=0, \quad \frac{d}{dt}(\gamma\dot{z})=0

Will come back later after finishing chapter 2

3.7: Motion in a Uniform Field

3.7.1: Doppler Effect

Dropper effect: the change in frequency of a wave when a source and an observer are moving relativity to each other. Frequency increases when the observer and the source approach each other and decreases when moving away.

Relativistic Dropper Shift

For a source directly approaching or moving away from an observer, the doppler effect gives:

fob=vv±vemfem,fob=fem11±vemvwavef_{ob}=\frac{v}{v\pm v_\text{em}}f_\text{em}, \quad f_{ob}=f_{em}\frac{1}{1\pm\frac{v_{em}}{v_{wave}}}
fob=fem1+v/c1v/cf_{ob}=f_{em}\sqrt{\frac{1+ v/c}{1-v/c}}

Redshift: As an object moves away from us, the sound or light waves emitted the source are stretched out, making them have a lower pitch and move towards the red end of the electromagnetic spectrum (which has a longer wavelength and a smaller frequency). The lights from the object is known as redshift.

Blueshift: As an object moves toward from us, the sound or light waves emitted the source shrink out, making them have a higher pitch and move towards the blue end of the electromagnetic spectrum (which has a shorted wavelength and a larger frequency). The lights from the object is known as blueshift.

3.7.2: Principle of Equivalence

Consider two spaceships. One accelerating uniformly at aa in gravity-free empty space and one standing at rest in a uniform gravity field. Let acceleration aa of the first ship is adjusted to be equal, but opposite in direction, to the gravitational field gg on the second ship.

An observer in the bow of the accelerating ship shies a laser beam at another observer in the stern of the ship. The ship is initially at rest and the laser emits monochromatic light of frequency femf_{em} in the laser’s rest frame. We assume that the distance traveled by the ship is very small while the bean is traveling (the assumption can be understood by considering the ship’s and the laser’s speed. The last has a speed of light CC, which we assume, is largely greater than the ship’s speed. Thus, the distance traveled by the ship is negligible compare to the distance traveled by the laser).

Let the length of the ship be hh. The time it takes for beam to reach the stern is t=h/ct = h/c. During the time, the stern attains a velocity v=at=ah/cv=at=ah/c.

fob=fem1+v/c1v/c=fob=fem(1+vc)=fem(1+ahc2)f_{ob}=f_{em}\sqrt{\frac{1+ v/c}{1-v/c}}=\approx f_{ob}=f_{em}(1+\frac{v}{c})=f_{em}(1+\frac{ah}{c^2})

Let h=yh=y. Because a=ga=g, the equation becomes:

fob=fem(1+ghc2)f_{ob}=f_{em}(1+\frac{gh}{c^2})

We can gain several insights from the equation. First, the frequency at the bottom (stern) is increased by a factor of gh/c2gh/c^2. In a time of t=1st=1s, the bow emits t/period=tfemt/\text{period}=tf_{em} wavelengths of light. However, because fob>femf_{ob}>f_{em}, observers at the bottom must collect theses waves in less than 1 second according to their own clocks (number of wavelengths of light remains unchanged for observers at the bottom and at the top, frequency fob>femf_{ob}>f_{em}, so time decreases).

tfem=n,n=number of wavelengthstf_{em}=n, \quad n=\text{number of wavelengths}
tem=nfem,tob=nfob=nfem(1+gy/c2)=tem1+gy/c2t_{em}=\frac{n}{f_{em}},\quad t_{ob}=\frac{n}{f_{ob}}=\frac{n}{f_{em}(1+gy/c^2)}=\frac{t_{em}}{1+gy/c^2}

Becase tlow=tob<tem=thight_{low}=t_{ob}<t_{em}=t_{high}, it is natural for observers at the top to conclude that bottom clocks run slower than top clocks (as their clocks have a smaller reading). The must be true for all clocks at the top and bottom. Rearranging the equation, we get

Δthigh=Δtlow(1+gy/c2)\boxed{\Delta t_{high}=\Delta t_{low}(1+gy/c^2)}

The equation shows the time difference for two clocks at rest, but at different altitudes in a uniform gravitational field.

Clock CC travels with a nonzero speed by an arbitrary path between stationary ground ground clocks AA and BB. All clocks are synchronized initially. Clock BB reads tft_f when CC arrives, and the reading of clock CC depends on the path it takes in moving from AA to BB.

Recall that the proper time for time dilation is:

τ=Δt1v2c2\tau =\Delta t\sqrt{1-\frac{v^2}{c^2}}

where vv is the velocity of the moving object relative to the outside observer. Because clock CC moves with a non-zero speed, by time dilation, we know that the reading of CC is different from the reading of BB. In an infinitesimal time dtdt according to the top clock, clock CC advances by the proper time:

dτ=dt(1+gy/c2)1v2/c2dt(1+gy/c2v2/2c2)d\tau=dt(1+gy/c^2)\sqrt{1-v^2/c^2}\approx dt(1+gy/c^2-v^2/2c^2)
τf=0tfdt(1+gy/c2v2/2c2)\boxed{\tau_f=\int^{t_f}_0 dt(1+gy/c^2-v^2/2c^2)}

3.7.3: The Hafele–Keating Experiments

Time for eastward clock
Westward clock

Explanation

The opposite sign of the two results suggests that the rate at which time flows depends on the motion of the observer. The eastward clock is moving in the same direction as earth’s rotation, so its velocity relative to earth is greater than the westward clock. Thus, a greater time dilation effect. The westward clock is moving in the opposite direction as the earth’s rotation, so it has a smaller relative velocity and time dilation effect.

Previously we have derived that clock CC advances by the proper time

τf=0tfdt(1+gy/c2v2/2c2)\tau_f=\int^{t_f}_0 dt(1+gy/c^2-v^2/2c^2)

We can rewrite the integral in the form of energy, giving us:

τf=0tfdt+(gy/c2v2/2c2)dt=tf+(v2/2c2+gy/c2)dt=tf1mc2(12mv2mgy)dt\begin{align*} \tau_f &=\int^{t_f}_0 dt+ \int (gy/c^2-v^2/2c^2)dt\\ &=t_f+\int (-v^2/2c^2+gy/c^2)dt\\ &=t_f-\frac{1}{mc^2}\int(\frac{1}{2}mv^2-mgy)dt \end{align*}

We know that 1/2mv2=K1/2mv^2=K is kinetic energy and mgy=Umgy=U is potential energy. Then, the integral becomes:

τf=tf0tf(KU)dt\begin{equation} \tau_f=t_f-\int^{t_f}_0(K-U)dt \end{equation}

In order to find a path that extremeizes (We use the term “extremeize” because it is uncertain if the proper time τ\tau is maximized or minimized. Euler’s equation only provides a path (a function) that maximizes or minimizes a function) the proper time τ\tau, we apply Euler’s equation to the integral II, giving us:

I=0tf(KU)dtI=\int^{t_f}_0(K-U)dt

Let F=F(x˙,x,y˙,y,z˙,z)F=F(\dot{x}, x, \dot{y}, y, \dot{z}, z ) be the integrand, we get:

F=F(x˙,x,y˙,y,z˙,z)=12m(x˙2+y˙2+z˙2)mgyF=F(\dot{x}, x, \dot{y}, y, \dot{z}, z )=\frac{1}{2}m(\dot{x}^2+\dot{y}^2+\dot{z}^2)-mgy

Since FF is independent of xx and zz, we get:

0=ddtFx,0=mgddtFy,0=ddtFz0=\frac{d}{dt}\frac{\partial F}{\partial x'},\quad 0=mg-\frac{d}{dt}\frac{\partial F}{\partial y'},\quad 0=\frac{d}{dt}\frac{\partial F}{\partial z'}\quad

giving us:

0=x˙x¨,0=gy˙y¨,0=z˙z¨0=\dot{x}\ddot{x},\quad 0=g-\dot{y}\ddot{y}, \quad 0=\dot{z}\ddot{z}

Then,

0=x¨,y˙=g,0=z¨0=\ddot{x}, \quad \dot{y}=g,\quad 0=\ddot{z}

When x¨,z¨=0\ddot{x}, \ddot{z}=0, we recognize these differential equations as results of Newton’s equation of motion F=maF=ma in a uniform gravitational field.

x=vxt,y=12gt2,z=0x=v_xt,\quad y=\frac{1}{2}gt^2, \quad z=0

Our goal in identifying a variational principle for the motion of a particle in a uniform gravitational field is successful. Without every using Newton’s law, we found the correct equations of motion. In the next section, we will derive F=UF=-\nabla U using variational calculus.

3.8: Arbitrary Potential Energy

I=0tf(KU)dtI=\int^{t_f}_0(K-U)dt

Let FF be the integrand, we get:

F(x˙,x,y˙,y,z˙,z)=12m(x˙2+y˙+z˙2)U(x,y,z)F(\dot{x}, x, \dot{y}, y, \dot{z}, z)=\frac{1}{2}m(\dot{x}^2+\dot{y}^+\dot{z}^2)-U(x,y,z)

Using Euler’s equation, we get:

0=ddtFx,0=mgddtFy,0=ddtFz0=\frac{d}{dt}\frac{\partial F}{\partial x'},\quad 0=mg-\frac{d}{dt}\frac{\partial F}{\partial y'},\quad 0=\frac{d}{dt}\frac{\partial F}{\partial z'}\quad

Summary:

Spherical:ds2=r2+r2dθ2+r2sin2θϕ2\text{Spherical:}\quad ds^2=r^2+r^2d\theta^2+r^2\sin^2\theta\phi^2
Cylinderical:ds=r2dθ2+z2\text{Cylinderical:}\quad ds=\sqrt{r^2d\theta^2+z^2}
t=dt=tdC/n(r)=1Cdsn(r)t=\int dt=t\int\frac{d}{C/n(r)}=\frac{1}{C}\int \frac{ds}{n(r)}
0=Fy(x)ddx(Fy(x))0=\frac{\partial F}{\partial y(x)}-\frac{d}{dx}\left(\frac{\partial F}{\partial y'(x)}\right)
Fyi(x)ddx(Fyi(x))=0\frac{\partial F}{\partial y_i(x)}-\frac{d}{dx}\left(\frac{\partial F}{\partial y_i'(x)}\right)=0
Δthigh=Δtlow(1+gy/c2)\Delta t_{high}=\Delta t_{low}(1+gy/c^2)
dτ=dt(1+gyc2v22c2)d\tau=dt(1+\frac{gy}{c^2}-\frac{v^2}{2c^2})
τf=tf0tf(KU)dt\tau_f=t_f-\int^{t_f}_0(K-U)dt

Practice Problems