5 Relativity

5.1 Galilean Relativity and the Death of Ether

Special relativity did not fall from the sky. It grew out of the success and then the failure of an older framework: Galilean relativity. In this section we build that framework, push it against electromagnetism, watch it crack, and see why the ether had to go.


5.1.1 Galileo’s ship: the original relativity principle

Galileo asked us to imagine a ship gliding smoothly on calm seas. Below deck, flies buzz, droplets fall, and a tossed key lands in your hand as if the ship were at rest. From these observations he distilled a principle:

Galilean relativity: The laws of mechanics have the same form in any inertial frame moving at constant velocity relative to another.


5.1.2 Mathematics of Galilean transformations

Let two inertial frames S and S' be related by a uniform relative velocity v along the x-axis. The transformation is

t=t,x=xvt,y=y,z=z\begin{aligned} t' &= t,\\ x' &= x - v t,\\ y' &= y,\\ z' &= z \end{aligned}

Velocities transform as

ux=uxv,uy=uy,uz=uz\begin{aligned} u'_x &= u_x - v,\\ u'_y &= u_y,\\ u'_z &= u_z \end{aligned}

Accelerations are invariant

a=a\boldsymbol a' = \boldsymbol a

A famous corollary is the velocity addition rule

ux=ux+vu_x = u'_x + v

Because acceleration is the same in all inertial frames, Newton’s second law ma=Fm\boldsymbol a=\boldsymbol F keeps its form so long as forces depend on relative positions and velocities in a way compatible with these rules.


5.1.3 Where waves crash into Galileo

Consider a one-dimensional wave equation

2ϕx21c22ϕt2=0\frac{\partial^2 \phi}{\partial x^2} - \frac{1}{c^2}\,\frac{\partial^2 \phi}{\partial t^2} = 0

Under a Galilean transformation x=xvt,  t=tx' = x - v t,\; t' = t, the derivatives mix and the equation in S' becomes

2ϕx21c2(t+vx) ⁣2ϕ=0\frac{\partial^2 \phi}{\partial x'^2} - \frac{1}{c^2}\left(\frac{\partial}{\partial t'} + v\,\frac{\partial}{\partial x'}\right)^{\!2}\phi = 0

which is not the same form. In Galilean kinematics, a wave equation implicitly names a preferred medium: the wave’s speed is measured relative to that stuff. For sound, the medium is air, and a ground observer measures approximately csound±vc_{\text{sound}}\pm v.

Maxwell’s equations in vacuum predict electromagnetic waves with speed c but introduce no material medium. Trying to force Maxwell into a Galilean world suggests different observers should measure c±vc\pm v, which experiments did not support.


5.1.4 The ether hypothesis

Nineteenth-century physicists proposed a medium so subtle that it filled space without dragging planets: the luminiferous ether. Light would be a transverse vibration of this ether, just as sound is a vibration of air, so c would be the speed relative to the ether and ordinary Galilean addition would apply.

Fizeau’s 1851 running-water experiment sent light through moving water and found that the light speed was not simply added as c/n±vc/n \pm v but modified by a Fresnel drag factor. A compact way to encode the observed effective speed was

vlight in watercn+v(11n2)v_{\text{light in water}} \approx \frac{c}{n} + v\left(1 - \frac{1}{n^2}\right)

People interpreted this as partial entrainment of ether by matter. The ether survived—battered, patched, never directly observed.


5.1.5 Michelson–Morley: the null result that echoed

If Earth moves through a stationary ether, there should be an ether wind. Light along or against that wind would have slightly different travel times compared with light across the wind. The Michelson–Morley interferometer (1887) split a beam, sent parts along perpendicular arms, recombined them, and watched for fringe shifts as the instrument rotated. They found essentially no shift. Repetitions at different times of year again gave null results. The expected v2/c2v^2/c^2 effect at orbital speeds did not appear.

Other experiments piled on: Trouton–Noble attempted to detect a torque on a moving charged capacitor; Kennedy–Thorndike examined frequency stability in a different geometry. Null, null. If an ether exists, it is undetectable not just dynamically but kinematically.


5.1.6 Lorentz’s rescue attempts: contraction and local time

Before Einstein, the best minds tried to save the ether by modifying matter, not spacetime. George FitzGerald and Hendrik Lorentz proposed that objects contract along the direction of motion by a factor

1v2c2\sqrt{1 - \frac{v^2}{c^2}}

Lorentz went further. To preserve the equations of electromagnetism, he introduced a mathematical device he called local time, and then discovered the full transformation that keeps Maxwell’s equations invariant

t=γ ⁣(tvxc2),x=γ(xvt),y=y,z=z,γ=11v2/c2\begin{aligned} t' &= \gamma\!\left(t - \frac{v x}{c^2}\right),\\ x' &= \gamma (x - v t),\\ y' &= y,\\ z' &= z,\\ \gamma &= \frac{1}{\sqrt{1 - v^2/c^2}} \end{aligned}

Lorentz still imagined a real ether and treated these relations as a masking effect that hid it. Mathematically, Lorentz had the right symmetry. Conceptually, the ether had become a spectator with no measurable role.


5.1.7 Why Galilean relativity fails for light

We can now state the clash cleanly. Galilean kinematics assumes absolute time and simple velocity addition. Maxwell’s electrodynamics predicts a universal wave speed c in vacuum. Experiments do not see c±vc\pm v or any ether wind. Keeping the Galilean structure forces a conspiracy of length contraction, clock distortions, and dynamical forces to cancel every observable effect, for all materials, orientations, seasons, and devices, which is implausible. Alternatively, keep c and the null results and replace the kinematics with one that respects a universal c.


5.1.8 Anatomy of the failure: a worked contrast

Let a light pulse move along +x in frame S

x=ctx = c t

Under a Galilean transformation with x=xvt,  t=tx' = x - v t,\; t' = t, the same worldline becomes

x=(cv)tx' = (c - v) t'

so the measured speed would be cvc - v. That is the collision with Maxwell.

By contrast, under a Lorentz transformation the same worldline remains lightlike and satisfies x=ctx' = c t' for every inertial observer because Lorentz transformations are precisely the linear maps that keep the light cone unchanged.


5.1.9 What about Doppler, aberration, and Fizeau?

Three often-confused phenomena align with relativity when analyzed carefully.

Doppler shift is a change in observed frequency due to relative motion; its relativistic formula reduces to the classical one at low speeds and matches high-speed astronomical data.

Aberration of starlight is the apparent tilt of stars due to finite light speed and Earth’s motion; it is consistent with relativity and does not require an ether wind inside instruments.

Fizeau’s running-water result emerges from relativistic velocity addition for light in a moving medium of refractive index nn and needs no ether substance.


5.1.10 Conceptual audit: what survives from Galileo?

Galileo’s core idea—that physics is the same in all inertial frames—survives. What changes is the implementation. In Galilean kinematics, time is absolute and the transformation group preserves tt. In relativistic kinematics, the invariant is the spacetime interval

Δs2=c2Δt2Δx2Δy2Δz2\Delta s^2 = c^2 \Delta t^2 - \Delta x^2 - \Delta y^2 - \Delta z^2

and the transformation group (the Lorentz group) preserves this quantity.


5.1.11 The death of ether (and what replaced it)

By the early twentieth century, the ether had been reduced to a metaphysical scaffold that explained nothing and predicted nothing new. Einstein’s 1905 paper did something radical by doing something minimal: he discarded the ether and kept only two principles—relativity and constant c—then showed that Lorentz’s mathematics flows from them. Space and time became spacetime; electric and magnetic fields became aspects of one tensor; simultaneity lost its absolute status. What replaced the ether was not a new material but a new geometry. The structure carrying electromagnetic waves is the metric of spacetime itself.


5.1.12 Takeaways

  • Galilean transformations encode absolute time and simple velocity addition; they keep Newton’s laws invariant but fail for Maxwell’s equations
  • The ether hypothesis tried to reconcile a universal c with Galilean kinematics and collapsed under precise null experiments
  • Lorentz found the right symmetry mathematically; Einstein made it the physics, elevating the invariance of c to a principle and retiring the ether
  • What remains of Galileo is the relativity principle itself, now implemented by the Lorentz group and an invariant spacetime interval

5.2 Special Relativity: Postulates and Kinematics

Special relativity is built on two simple principles and a lot of courage. The principles are

(i) the relativity principle—the laws of physics have the same form in all inertial frames

(ii) the constancy of the speed of light—every inertial observer measures light in vacuum to move at speed cc,

regardless of the motion of source or observer. From these, the familiar structure of spacetime follows.


5.2.1 Synchronizing clocks and the fall of simultaneity

Einstein proposed a physical protocol to define simultaneity. To synchronize distant clocks, send a light pulse from A to B and back, and set B so the outbound and inbound legs take equal times. This construction is frame-dependent because different inertial observers slice spacetime differently. Thus, simultaneity is not absolute; it depends on motion. Timekeeping becomes geometry.


5.2.2 Deriving the Lorentz transformation

Consider two frames SS and SS' in standard configuration: SS' moves with constant velocity vv along the xx-axis of SS, and origins coincide at t=t=0t=t'=0. Homogeneity and isotropy of spacetime imply a linear relation between coordinates. Requiring that the worldline of a light pulse along +x+x be x=ctx=ct in SS and x=ctx'=ct' in SS' fixes the transformation up to a scale, which normalization at v=0v=0 removes. The result is

x=γ(xvt),t=γ ⁣(tvxc2),y=y,z=z,γ=11v2/c2\begin{aligned} x' &= \gamma (x - v t),\\ t' &= \gamma \!\left(t - \frac{v x}{c^{2}}\right),\\ y' &= y,\qquad z' = z,\\ \gamma &= \frac{1}{\sqrt{1 - v^{2}/c^{2}}} \end{aligned}

The inverse transformation (swap vvv\to -v) is

x=γ(x+vt),t=γ ⁣(t+vxc2),y=y,z=z\begin{aligned} x &= \gamma (x' + v t'),\\ t &= \gamma \!\left(t' + \frac{v x'}{c^{2}}\right),\\ y &= y',\qquad z = z' \end{aligned}

These are the unique linear maps that preserve light cones.


5.2.3 Invariant interval and proper time

From the Lorentz transformation one finds the Lorentz-invariant spacetime interval

Δs2=c2Δt2Δx2Δy2Δz2\Delta s^{2} = c^{2}\Delta t^{2} - \Delta x^{2} - \Delta y^{2} - \Delta z^{2}

For a timelike worldline, define the proper time dτd\tau via c2dτ2=ds2c^{2} d\tau^{2} = ds^{2}. Integrating along a path yields the time experienced by the moving clock. Because ds2ds^{2} is invariant, all inertial observers agree on τ\tau for the same worldline.


5.2.4 Time dilation, length contraction, simultaneity shift

A light clock carried by a moving observer, or directly from the transformation, gives

Δt=γΔτ\Delta t = \gamma\, \Delta \tau

A moving clock runs slow by a factor γ\gamma. For a rod of rest length L0L_{0} aligned with the xx-axis, measuring its endpoints simultaneously in SS yields

L=L0γL = \frac{L_{0}}{\gamma}

Simultaneity is frame-dependent. Two events separated by (Δt,Δx)(\Delta t,\,\Delta x) in SS have time separation in SS' given by

Δt=γ ⁣(ΔtvΔxc2)\Delta t' = \gamma \!\left(\Delta t - \frac{v\,\Delta x}{c^{2}}\right)

Even if Δt=0\Delta t=0 in SS (simultaneous there), generally Δt0\Delta t'\neq 0 in SS'. That is the operational content of the relativity of simultaneity.


5.2.5 Velocity addition and transverse components

Velocities do not add linearly. Differentiating the Lorentz transformation gives the Einstein velocity addition rules. For a particle with velocity components (ux,uy,uz)(u_x,u_y,u_z) in SS, the velocity seen in SS' is

ux=uxv1vuxc2u'_x = \frac{u_x - v}{1 - \dfrac{v\,u_x}{c^{2}}} uy=uyγ ⁣(1vuxc2),uz=uzγ ⁣(1vuxc2)u'_y = \frac{u_y}{\gamma \!\left(1 - \dfrac{v\,u_x}{c^{2}}\right)},\qquad u'_z = \frac{u_z}{\gamma \!\left(1 - \dfrac{v\,u_x}{c^{2}}\right)}

For collinear motion, the composition law reduces to

uv=u+v1+uvc2u \oplus v = \frac{u + v}{1 + \dfrac{u v}{c^{2}}}

No matter how you stack subluminal speeds, the result remains <c<c.

It is often convenient to use rapidity η\eta defined by βv/c=tanhη\beta \equiv v/c = \tanh \eta. Then

γ=coshη,γβ=sinhη\gamma = \cosh \eta,\qquad \gamma\beta = \sinh \eta

and collinear velocities add by adding rapidities,

ηtotal=η1+η2\eta_{\text{total}} = \eta_{1} + \eta_{2}

This turns kinematics into simple hyperbolic geometry.


5.2.6 Geometry of Minkowski diagrams

Events are points; worldlines are curves. The light cone at any event divides spacetime into timelike future/past (inside the cone), spacelike elsewhere (outside), and lightlike (on the cone). Causality requires signals to stay inside or on light cones. Lorentz transformations act as hyperbolic rotations that keep cones fixed and shear constant-time slices; that is why simultaneity changes but light speed does not.


5.2.7 Standard “paradoxes” and how the math resolves them

Twin scenario. One twin travels fast and returns younger. The traveling twin’s worldline has smaller proper time because it changes inertial frames (nonzero acceleration at turn-around). Proper time is the integral of dτd\tau along the path, not a naive “time runs slower over there” slogan.

Barn–ladder scenario. A fast ladder fits into a short barn due to L=L0/γL=L_0/\gamma in the barn frame. In the ladder’s frame, the doors are not closed simultaneously; the relativity of simultaneity saves causality. The transformation formula for Δt\Delta t' above is the whole story.

Magnet and conductor. A moving charge near a neutral wire experiences an electric field in its frame and a magnetic force in the lab frame; both arise from one electromagnetic tensor under Lorentz transformation. Kinematics and fields are consistent.


5.2.8 Useful algebraic identities

The Lorentz factor satisfies

γ2(1β2)=1,γ2(γβ)2=1\gamma^{2} (1 - \beta^{2}) = 1,\qquad \gamma^{2} - (\gamma\beta)^{2} = 1

For small speeds vcv\ll c, expand

γ1+12β2+38β4+\gamma \approx 1 + \tfrac{1}{2}\beta^{2} + \tfrac{3}{8}\beta^{4} + \cdots

Relativistic Doppler for motion directly along the line of sight uses the Doppler factor

D=1+β1β\mathcal{D} = \sqrt{\frac{1+\beta}{1-\beta}}

so observed frequency satisfies νobs=Dνemit\nu_{\text{obs}}=\mathcal{D}\,\nu_{\text{emit}} for approaching motion and νobs=νemit/D\nu_{\text{obs}}=\nu_{\text{emit}}/\mathcal{D} for receding motion.


5.2.9 Takeaways

  • The Lorentz transformation is the unique linear mapping that preserves the light cone and follows from the two postulates.
  • Time dilation, length contraction, and simultaneity shift are different faces of the same transformation.
  • Velocity addition prevents superluminal composition; rapidity linearizes collinear boosts.
  • The invariant interval underlies proper time and causality; Minkowski geometry is the natural language of kinematics.

Next we will recast dynamics—momentum, energy, and E=mc2E=m c^{2}—in this geometric framework.


5.3 Minkowski Spacetime and Causality

Special relativity turns kinematics into geometry. Space and time do not live in separate ledgers; they form a single four-dimensional arena—Minkowski spacetime—whose geometry encodes causality. In this section we define the metric, classify separations as timelike/spacelike/lightlike, introduce four-vectors and boosts as hyperbolic rotations, and extract practical rules for reading spacetime diagrams.


5.3.1 Events, coordinates, and the metric

An event is a point in spacetime with coordinates xμ=(ct,x,y,z)x^\mu = (ct,\,x,\,y,\,z). Latin letters will denote spatial vectors x=(x,y,z)\mathbf x=(x,y,z); Greek indices run over 0,1,2,30,1,2,3. In inertial Cartesian coordinates the Minkowski metric is

ημν=diag(1,1,1,1)\eta_{\mu\nu}=\mathrm{diag}(1,\,-1,\,-1,\,-1)

The invariant squared interval between two nearby events xμx^\mu and xμ+dxμx^\mu+dx^\mu is

ds2=ημνdxμdxν=c2dt2dx2ds^2 = \eta_{\mu\nu}\,dx^\mu dx^\nu = c^2 dt^2 - d\mathbf x^2

All inertial observers compute the same ds2ds^2. This single scalar encodes what is physically meaningful about separations.


5.3.2 Light cones and the three kinds of separation

For any pair of events A and B, the sign of s2s^2 classifies their relation:

Light cone

  • Timelike: s2>0s^2>0. There exists a frame in which the events occur at the same spatial point but different times; a slower-than-light traveler can go from A to B.
  • Lightlike (null): s2=0s^2=0. Only a light signal can connect them.
  • Spacelike: s2<0s^2<0. There exists a frame in which the events are simultaneous but different places; no signal can connect them without exceeding cc.

At any event the set ds2=0ds^2=0 forms the light cone with generators x=±ctx=\pm ct (in 1+11+1 dimensions). The interior is timelike, the exterior spacelike. Lorentz transformations preserve this cone, so all inertial observers agree on causal possibilities even if they disagree on times and distances.


5.3.3 Proper time and worldlines

A worldline is a curve xμ(λ)x^\mu(\lambda) describing an object’s history. For timelike motion the clock carried by the object ticks the proper time

dτ=1cds2=dt2dx2c2d\tau = \frac{1}{c}\,\sqrt{ds^2} = \sqrt{dt^2 - \frac{d\mathbf x^{\,2}}{c^2}}

Integrating gives the elapsed time on that clock:

Δτ=pathdτ\Delta \tau = \int_{\text{path}} d\tau

Among all timelike curves connecting the same endpoints in flat spacetime, inertial motion maximizes Δτ\Delta\tau. This geometric statement resolves the twin scenario: the traveling twin takes a non-inertial path with smaller proper time.


5.3.4 Four-velocity and four-acceleration

Define the four-velocity

uμdxμdτu^\mu \equiv \frac{dx^\mu}{d\tau}

Using ds2=c2dτ2ds^2=c^2 d\tau^2 one finds the normalization

ημνuμuν=c2\eta_{\mu\nu} u^\mu u^\nu = c^2

Writing spatial velocity v=dx/dt\mathbf v=d\mathbf x/dt and γ=1/1v2/c2\gamma=1/\sqrt{1-v^2/c^2} gives

uμ=γ(c,v)u^\mu = \gamma(c,\,\mathbf v)

The four-acceleration aμ=duμ/dτa^\mu = du^\mu/d\tau satisfies an orthogonality identity

ημνuμaν=0\eta_{\mu\nu} u^\mu a^\nu = 0

meaning acceleration is purely “spacelike” in the instantaneous rest frame. These tools streamline relativistic dynamics (§5.5).


5.3.5 Lorentz transformations as hyperbolic rotations

A boost along xx with speed vv can be written using β=v/c\beta=v/c and γ=1/1β2\gamma=1/\sqrt{1-\beta^2}:

ct=γ(ctβx)x=γ(xβct)y=yz=z\begin{aligned} ct' &= \gamma\,(ct - \beta x)\\ x' &= \gamma\,(x - \beta ct)\\ y' &= y\\ z' &= z \end{aligned}

The matrix Λμν\Lambda^\mu{}_\nu of this transformation obeys the invariance condition

ηαβΛαμΛβν=ημν\eta_{\alpha\beta}\,\Lambda^\alpha{}_\mu\,\Lambda^\beta{}_\nu = \eta_{\mu\nu}

This is analogous to how ordinary rotations preserve Euclidean lengths. Indeed, if we define the rapidity η\eta by β=tanhη\beta=\tanh\eta, then

γ=coshη,γβ=sinhη\gamma=\cosh\eta,\qquad \gamma\beta=\sinh\eta

and the boost becomes a hyperbolic rotation in the ctctxx plane. Composition of collinear boosts simply adds rapidities. That is why velocity composition has the non-linear Einstein form yet remains associative in rapidity space.


5.3.6 Reading Minkowski diagrams

A Minkowski diagram displays ctct vertically and xx horizontally. Light rays draw lines at slopes ±45\pm 45^\circ (since x=±ctx=\pm ct in units with c=1c=1). Some practical rules:

Minkowski diagram

  1. Axes of a moving frame. The xx' axis is the line of events with t=0t'=0, which tilts upward; the ctct' axis is the worldline x=0x'=0, which tilts right. The two axes are symmetric with respect to the light cone.
  2. Simultaneity slices. Lines of constant tt' are not horizontal unless v=0v=0; that is the relativity of simultaneity made visible.
  3. Unit calibration. The unit hyperbola ct=x2+c2τ2ct=\sqrt{x^2+c^2\tau^2} encodes proper time ticks; Loedel diagrams use symmetric scaling to keep both frames drawn with the same light-cone angle.
  4. Causality test. If an event B lies outside A’s light cone, no worldline with v<c|v|<c connects A to B. Event order for spacelike pairs can flip under boosts; for timelike pairs it cannot.

Diagrams are not just pictures; they are calculators for ordering, time dilation, length contraction, and simultaneity shift.


5.3.7 Causality, no-signaling, and why cc is a speed limit

Suppose there existed a signal with speed u>cu>c in some frame. A boost can be found that flips the temporal order of emission and reception for a spacelike separation. Combined with “you can send signals both ways,” this makes a closed causal loop—a paradox. Special relativity avoids this by forbidding superluminal signaling: physical worldlines are timelike or null, never spacelike. As a corollary, there are no perfectly rigid bodies; attempts to jerk one end of a rod cannot simultaneously affect the other end because stress information travels at material sound speeds <c<c.

Microcausality in field theory is the quantum echo of this rule: field operators at spacelike separations commute so that measurements cannot influence one another.


5.3.8 Energy–momentum as a four-vector (preview)

Later we will use the energy–momentum four-vector

pμ=(Ec,p)p^\mu = \left(\frac{E}{c},\,\boldsymbol p\right)

with invariant

pμpμ=(Ec)2p2=m2c2p_\mu p^\mu = \left(\frac{E}{c}\right)^2 - \boldsymbol p^{\,2} = m^2 c^2

This single relation encodes both the rest-energy E0=mc2E_0=mc^2 and the dispersion E2=(pc)2+(mc2)2E^2=(pc)^2+(mc^2)^2. The conservation of pμp^\mu across interactions is a compact, frame-independent way to do dynamics.


5.3.9 Accelerated motion in flat spacetime

Even without gravity we can treat accelerated observers. A worldline with constant proper acceleration aa in the xx direction traces a hyperbola,

ct(τ)=c2asinh ⁣(aτc)x(τ)=c2acosh ⁣(aτc)\begin{aligned} ct(\tau) &= \frac{c^2}{a}\,\sinh\!\left(\frac{a\tau}{c}\right)\\ x(\tau) &= \frac{c^2}{a}\,\cosh\!\left(\frac{a\tau}{c}\right) \end{aligned}

which stays asymptotic to the light lines x=±ctx=\pm ct. Proper acceleration is what an accelerometer reads; it is invariant, unlike coordinate acceleration which depends on frame.

Accelerated frames bring in Rindler coordinates and horizons even in flat spacetime—useful stepping stones toward the equivalence principle (§5.8).


5.3.10 Tensors and index gymnastics (minimal toolkit)

A four-vector transforms as Aμ=ΛμνAνA'^\mu=\Lambda^\mu{}_\nu A^\nu. Lowering and raising indices uses ημν\eta_{\mu\nu} and its inverse:

Aμ=ημνAν,Aμ=ημνAνA_\mu = \eta_{\mu\nu} A^\nu,\qquad A^\mu = \eta^{\mu\nu} A_\nu

Lorentz-invariant scalars are built by contracting indices, for example AμBμA_\mu B^\mu. The electromagnetic field appears as an antisymmetric tensor FμνF_{\mu\nu}; writing Maxwell’s equations in this language exposes symmetry and automatically preserves causality, because FμνF_{\mu\nu} transformations keep light cones intact.


5.3.11 What changes, what stays the same

  • Changes: time intervals, lengths, and simultaneity slices between frames.
  • Stays the same: the light cone, the interval ds2ds^2, proper time τ\tau, and causal order for timelike-related events.

This is the mantra of relativistic reasoning: compute invariants first; everything else is perspective.


5.3.12 Quick problem templates

  1. Is A able to influence B? Compute s2s^2. If s2<0s^2<0, no. If s2=0s^2=0, only with light. If s2>0s^2>0, yes with subluminal motion.
  2. How old is the traveler? Integrate dτ=1v2/c2dtd\tau=\sqrt{1-v^2/c^2}\,dt along the actual velocity profile.
  3. What does a moving observer see? Apply the boost; redraw axes; use the simultaneity formula and length contraction as needed.

In summary: Minkowski spacetime packages relativity into geometry. The metric ds2=c2dt2dx2ds^2=c^2dt^2-d\mathbf x^2 defines light cones and proper time; boosts are hyperbolic rotations preserving those cones. Causality becomes a geometric constraint—worldlines must stay inside or on light cones—and dynamics a story of four-vectors that transform covariantly. Once you think in intervals, cones, and rapidities, the paradoxes evaporate and the calculations streamline.


5.4 Relativistic Dynamics

Kinematics told us how coordinates transform. Dynamics asks how momentum, energy, and forces behave so that the laws look the same to every inertial observer. The guiding idea is simple: build everything from Lorentz-invariant or covariant objects so that physical predictions do not depend on the frame.


5.4.1 Four-momentum and the energy–momentum relation

Define the four-velocity and four-momentum

uμdxμdτ,pμmuμu^\mu \equiv \frac{dx^\mu}{d\tau}, \qquad p^\mu \equiv m u^\mu

with dτd\tau the proper time. Using uμuμ=c2u^\mu u_\mu = c^2, we have

pμ=γ(mc,mv),γ11v2/c2p^\mu = \gamma \,(m c,\, m \mathbf v), \qquad \gamma \equiv \frac{1}{\sqrt{1 - v^2/c^2}}

The temporal component defines energy

Ep0c=γmc2E \equiv p^0 c = \gamma m c^2

and spatial part defines momentum

pγmv\boldsymbol p \equiv \gamma m \boldsymbol v

The invariant magnitude of pμp^\mu gives the master relation

pμpμ=m2c2p_\mu p^\mu = m^2 c^2

or, written in scalars,

E2=(pc)2+(mc2)2E^2 = (p c)^2 + (m c^2)^2

This reduces at low speeds to familiar Newtonian forms. Expanding γ\gamma for vcv \ll c,

E=γmc2mc2+p22m+E = \gamma m c^2 \approx m c^2 + \tfrac{p^2}{2m} + \cdots

so the Newtonian kinetic energy appears as the first correction above the rest energy mc2m c^2.


5.4.2 Work, power, and the relativistic force law

Keep the Newtonian definitions but use relativistic p\boldsymbol p and EE:

Fdpdt,dEdt=Fv\boldsymbol F \equiv \frac{d\boldsymbol p}{dt}, \qquad \frac{dE}{dt} = \boldsymbol F \cdot \boldsymbol v

With p=γmv\boldsymbol p=\gamma m \boldsymbol v one finds

F=γma+γ3mv(v ⁣ ⁣a)c2\boldsymbol F = \gamma m \boldsymbol a + \gamma^3 m \frac{\boldsymbol v\,(\boldsymbol v\!\cdot\!\boldsymbol a)}{c^2}

Decomposing into parts parallel and perpendicular to v\mathbf v gives the compact rules

a=Fγ3m,a=Fγma_\parallel = \frac{F_\parallel}{\gamma^3 m}, \qquad a_\perp = \frac{F_\perp}{\gamma m}

So as speeds approach cc, the same applied force yields smaller acceleration, especially along the direction of motion. The speed limit is thus dynamical as well as kinematical.

It is often cleaner to package force into a four-vector. Define the four-force

KμdpμdτK^\mu \equiv \frac{dp^\mu}{d\tau}

which automatically satisfies an orthogonality identity

Kμuμ=0K^\mu u_\mu = 0

because pμpμp^\mu p_\mu is constant. In §5.6 we will specialize this to charged particles in electromagnetic fields.


5.4.3 Action principle and the free-particle Lagrangian

Relativistic dynamics drops out of a geometric action. For a neutral particle,

S=mcdsS = - m c \int ds

In a given frame, ds=cdt1v2/c2ds = c\,dt\,\sqrt{1 - v^2/c^2}, so the Lagrangian is

L0=mc21v2c2L_0 = - m c^2 \sqrt{1 - \frac{v^2}{c^2}}

The canonical momentum and Hamiltonian follow:

p=L0v=γmv\boldsymbol p = \frac{\partial L_0}{\partial \boldsymbol v} = \gamma m \boldsymbol v H=pvL0=γmc2=EH = \boldsymbol p \cdot \boldsymbol v - L_0 = \gamma m c^2 = E

Thus the energy–momentum relations are not assumptions but consequences of the variational principle built from the invariant length of worldlines.

When a particle carries charge qq and interacts with an electromagnetic four-potential Aμ=(ϕ/c,A)A_\mu=(\phi/c,\,-\boldsymbol A), the minimal-coupling action adds qAμdxμq\int A_\mu dx^\mu. In ordinary variables the Lagrangian becomes

L=mc21v2c2+qvAqϕL = - m c^2 \sqrt{1 - \frac{v^2}{c^2}} + q\,\boldsymbol v\cdot\boldsymbol A - q\,\phi

whose Euler–Lagrange equations yield the Lorentz force in covariant form (shown in §5.6).


5.4.4 Momentum and energy balance in collisions

Because four-momentum is conserved,

inpμ=outpμ\sum_{\text{in}} p^\mu = \sum_{\text{out}} p^\mu

This single vector equation encodes both ordinary momentum conservation and energy conservation in all frames. Squaring both sides produces frame-invariant relations useful for particle reactions.

A key invariant is the Mandelstam ss for two particles with four-momenta p1,p2p_1,p_2:

s(p1+p2)2s \equiv (p_1 + p_2)^2

In the center-of-momentum (CM) frame, s=(ECM/c)2s = (E_{\text{CM}}/c)^2, so ECM=scE_{\text{CM}} = \sqrt{s}\,c. In a lab frame where target bb is at rest and beam aa has energy EaE_a,

s=ma2c2+mb2c2+2mbEacs = m_a^2 c^2 + m_b^2 c^2 + 2 m_b \frac{E_a}{c}

or, more transparently,

sc2=ma2c4+mb2c4+2mbc2Eas c^2 = m_a^2 c^4 + m_b^2 c^4 + 2 m_b c^2 E_a

At threshold for producing final particles of total rest mass MfinalM_{\text{final}}, the CM kinetic energy just vanishes, so

sth=Mfinal2c2s_{\text{th}} = M_{\text{final}}^{\,2} c^2

and the required lab energy of the beam is

Ea,th=Mfinal2c4(ma2+mb2)c42mbc2E_{a,\text{th}} = \frac{M_{\text{final}}^{\,2} c^4 - \left(m_a^2 + m_b^2\right) c^4}{2 m_b c^2}

As a concrete example, consider p+pp+p+π0p + p \to p + p + \pi^0 with one proton at rest. Taking mpm_p and mπm_\pi as the rest masses, Mfinal=2mp+mπM_{\text{final}} = 2 m_p + m_\pi, and the formula above gives the familiar multi-GeV threshold for pion production in fixed-target experiments.


5.4.5 Center-of-momentum frame and invariant mass of a system

For any isolated system, define the total four-momentum

PμipiμP^\mu \equiv \sum_i p_i^\mu

The invariant

Msys2c2PμPμM_{\text{sys}}^2 c^2 \equiv P_\mu P^\mu

is the system’s invariant mass. In the CM frame P=0\boldsymbol P=\boldsymbol 0, so Msysc2=ECMM_{\text{sys}} c^2 = E_{\text{CM}}. Crucially, MsysM_{\text{sys}} is not just the sum of rest masses; it includes kinetic and binding energies. For example, a box containing radiation has larger MsysM_{\text{sys}} than the empty box, because photons carry energy and momentum even though their individual rest masses are zero.


5.4.6 Massless particles and dispersion

For m=0m=0 the invariant relation becomes

E=pcE = p c

The four-velocity is undefined for lightlike worldlines, but the four-momentum remains finite and transforms correctly. A photon’s energy relates to frequency by E=hνE=h\nu, consistent with p=h/λp=h/\lambda and Maxwell’s waves moving at cc. Massive particles cannot reach cc because that would require γ\gamma \to \infty, hence infinite energy.


5.4.7 Dynamics with constraints and radiation reaction (glimpse)

For charged particles, the Lorentz force law reads in three-vector form (derived covariantly in §5.6)

dpdt=q(E+v×B),dEdt=qEv\frac{d\boldsymbol p}{dt} = q\left(\boldsymbol E + \boldsymbol v \times \boldsymbol B\right), \qquad \frac{dE}{dt} = q\,\boldsymbol E\cdot\boldsymbol v

Because accelerating charges radiate, an additional radiation reaction force arises at higher fidelity. Its classical expression (Abraham–Lorentz–Dirac) is subtle and can lead to runaways; in practice one uses approximations consistent with the external fields. The key point for dynamics: energy lost to radiation must be accounted for in dE/dtdE/dt.


5.4.8 Proper acceleration and hyperbolic motion

A trajectory with constant proper acceleration aa along xx satisfies

ct(τ)=c2asinh ⁣(aτc)x(τ)=c2acosh ⁣(aτc)\begin{aligned} ct(\tau) &= \frac{c^2}{a}\,\sinh\!\left(\frac{a\tau}{c}\right)\\ x(\tau) &= \frac{c^2}{a}\,\cosh\!\left(\frac{a\tau}{c}\right) \end{aligned}

Differentiation gives speed

v(τ)=dxdt=ctanh ⁣(aτc)v(\tau) = \frac{dx}{dt} = c\,\tanh\!\left(\frac{a\tau}{c}\right)

So even with constant proper acceleration, the coordinate acceleration decreases as vv approaches cc, asymptotically respecting the speed limit while proper time accumulates at a slower rate.


5.4.9 Quick toolbox for problems

  • Use pμ=(E/c,p)p^\mu=(E/c,\,\boldsymbol p) and conserve four-momentum; square once if you want frame-free answers.
  • For rockets or beams, compute power with dE/dt=FvdE/dt=\boldsymbol F\cdot\boldsymbol v and relate thrust to dp/dtd\boldsymbol p/dt.
  • Choose the CM frame to minimize algebra; transform back with the Lorentz formulas of §5.3.
  • Remember MsysM_{\text{sys}} includes kinetic and binding energy; “mass of a system” is an invariant, not just the sum of constituent rest masses.

In summary: Relativistic dynamics packages energy and momentum into a single four-vector that obeys E2=(pc)2+(mc2)2E^2=(pc)^2+(mc^2)^2. Forces change momentum via dp/dtd\boldsymbol p/dt and do work at rate Fv\mathbf F\cdot\mathbf v, while the covariant four-force Kμ=dpμ/dτK^\mu=dp^\mu/d\tau keeps orthogonality to uμu^\mu. Collisions are clean in terms of invariants like ss and the system mass MsysM_{\text{sys}}. All of it is geometry: conserve four-vectors and respect the light cone, and dynamics becomes as straightforward as kinematics.


5.5 Electromagnetism in Relativistic Form

Maxwell’s theory already smells relativistic. Here we make that symmetry explicit: package E\boldsymbol E and B\boldsymbol B into a single field tensor, write the equations and forces in four-dimensional form, and expose the invariants that every observer agrees on.


5.5.1 Four-current, four-potential, and the field tensor

Combine charge density and current into the four-current

Jμ(cρ,J)J^\mu \equiv \left(c\rho,\,\boldsymbol J\right)

Combine scalar and vector potentials into the four-potential

Aμ(ϕc,A)A^\mu \equiv \left(\frac{\phi}{c},\,\boldsymbol A\right)

Define the antisymmetric electromagnetic field tensor

FμνμAννAμF_{\mu\nu} \equiv \partial_\mu A_\nu - \partial_\nu A_\mu

With metric ημν=diag(1,1,1,1)\eta_{\mu\nu}=\mathrm{diag}(1,-1,-1,-1) (SI units),

F0i=Eic,Fij=ϵijkBkF^{0i} = \frac{E_i}{c},\qquad F^{ij} = -\,\epsilon^{ijk} B_k

So E\boldsymbol E and B\boldsymbol B are not separate things; they are different facets of the same tensor.


5.5.2 Maxwell’s equations as two short tensor laws

All four Maxwell equations compress to just two:

μFμν=μ0Jν\partial_\mu F^{\mu\nu} = \mu_0 J^\nu [αFβγ]=0\partial_{[\alpha} F_{\beta\gamma]} = 0

Using the dual tensor F~μν12ϵμναβFαβ\tilde F^{\mu\nu} \equiv \tfrac12 \epsilon^{\mu\nu\alpha\beta} F_{\alpha\beta}, the second line becomes

μF~μν=0\partial_\mu \tilde F^{\mu\nu} = 0

Written this way, Lorentz covariance is manifest: the equations keep the same form in every inertial frame.


5.5.3 Lorentz force in one line

A particle of charge qq and four-velocity uμu^\mu obeys

dpμdτ=qFμ νuν\frac{dp^\mu}{d\tau} = q\,F^\mu{}_{\ \nu}\,u^\nu

Splitting into time and space parts gives the familiar three-vector form

dpdt=q(E+v×B),dEdt=qEv\frac{d\boldsymbol p}{dt} = q\left(\boldsymbol E + \boldsymbol v \times \boldsymbol B\right),\qquad \frac{dE}{dt} = q\,\boldsymbol E\cdot\boldsymbol v

One covariant equation, every frame happy.


5.5.4 Gauge symmetry and wave equations

Gauge transformations leave physics unchanged:

AμAμ+μχA_\mu \rightarrow A_\mu + \partial_\mu \chi

In the Lorenz gauge

μAμ=0\partial_\mu A^\mu = 0

Maxwell’s equations reduce to wave equations for the potential:

Aμ=μ0Jμ\square A^\mu = \mu_0 J^\mu

where μμ=1c22t22\square \equiv \partial_\mu\partial^\mu = \frac{1}{c^2}\frac{\partial^2}{\partial t^2}-\nabla^2 is the d’Alembertian.


5.5.5 How fields transform: boosts mix E\,\boldsymbol E and B\,\boldsymbol B

Relativity treats “electric” and “magnetic” as frame-dependent decompositions of one tensor. For a boost with velocity v\boldsymbol v,

E=E,B=B\boldsymbol E'_{\parallel} = \boldsymbol E_{\parallel},\qquad \boldsymbol B'_{\parallel} = \boldsymbol B_{\parallel} E=γ(E+v×B),B=γ(Bv×Ec2)\boldsymbol E'_{\perp} = \gamma\left(\boldsymbol E_{\perp} + \boldsymbol v \times \boldsymbol B\right),\qquad \boldsymbol B'_{\perp} = \gamma\left(\boldsymbol B_{\perp} - \frac{\boldsymbol v \times \boldsymbol E}{c^2}\right)

For a boost along xx with speed vv,

Ex=ExEy=γ ⁣(EyvBz)Ez=γ ⁣(Ez+vBy)\begin{aligned} E'_x &= E_x\\ E'_y &= \gamma\!\left(E_y - v B_z\right)\\ E'_z &= \gamma\!\left(E_z + v B_y\right) \end{aligned} Bx=BxBy=γ ⁣(By+vc2Ez)Bz=γ ⁣(Bzvc2Ey)\begin{aligned} B'_x &= B_x\\ B'_y &= \gamma\!\left(B_y + \frac{v}{c^2} E_z\right)\\ B'_z &= \gamma\!\left(B_z - \frac{v}{c^2} E_y\right) \end{aligned}

A pure electric field in one frame generally appears as a mix of electric and magnetic fields in another.


5.5.6 Two Lorentz invariants (frame-proof diagnostics)

Two independent Lorentz scalars diagnose the field:

I1FμνFμν\mathcal I_1 \equiv F_{\mu\nu}F^{\mu\nu} I212ϵμναβFμνFαβ\mathcal I_2 \equiv \tfrac12 \epsilon^{\mu\nu\alpha\beta}F_{\mu\nu}F_{\alpha\beta}

In SI variables,

I1=2 ⁣(B2E2c2),I2=4cEB\mathcal I_1 = 2\!\left(\boldsymbol B^2 - \frac{\boldsymbol E^2}{c^2}\right),\qquad \mathcal I_2 = -\,\frac{4}{c}\,\boldsymbol E\cdot\boldsymbol B

For a plane wave,

EB,E=cB,EB=0\boldsymbol E \perp \boldsymbol B,\qquad |\boldsymbol E| = c\,|\boldsymbol B|,\qquad \boldsymbol E\cdot\boldsymbol B = 0

so I1=I2=0\mathcal I_1=\mathcal I_2=0 (a null field). If I1<0\mathcal I_1<0 and I2=0\mathcal I_2=0, there exists a frame where the field is purely electric; if I1>0\mathcal I_1>0 and I2=0\mathcal I_2=0, a purely magnetic frame exists.


5.5.7 Field energy–momentum and stresses

Package energy, momentum, and stresses into the electromagnetic stress–energy tensor

Tμν=1μ0 ⁣(FμαFνα14ημνFαβFαβ)T^{\mu\nu} = \frac{1}{\mu_0}\!\left(F^{\mu\alpha}F^\nu{}_{\alpha} - \tfrac14 \eta^{\mu\nu} F^{\alpha\beta}F_{\alpha\beta}\right)

Its (non)conservation in the presence of sources reads

μTμν=FνλJλ\partial_\mu T^{\mu\nu} = -\,F^{\nu\lambda} J_\lambda

Components:

T00=12ε0E2+12μ0boldsymbolB2T^{00} = \frac{1}{2}\varepsilon_0 \boldsymbol E^2 + \frac{1}{2\mu_0}\\boldsymbol B^2 T0i=1c2Si,S=1μ0E×BT^{0i} = \frac{1}{c^2} S_i,\qquad \boldsymbol S = \frac{1}{\mu_0}\,\boldsymbol E \times \boldsymbol B

So T00T^{00} is energy density, and T0iT^{0i} encodes momentum density via the Poynting vector S\boldsymbol S. The spatial block TijT^{ij} (Maxwell stress tensor) gives pressures and tensions exerted by fields.


5.5.8 Lagrangian density (one line to rule them all)

The field-theory origin story fits on one line:

L=14μ0FμνFμνJμAμ\mathcal L = -\,\frac{1}{4\mu_0}\,F^{\mu\nu}F_{\mu\nu} - J^\mu A_\mu

Varying AμA_\mu yields Maxwell’s equations; gauge symmetry leads, via Noether’s theorem, to charge conservation μJμ=0\partial_\mu J^\mu=0.


5.5.9 Fields of moving charges (why magnetism appears)

Boost a static Coulomb field and you get an anisotropically compressed electric field plus a magnetic field. In other words,

magnetic field=“electric field seen from a moving frame”\text{magnetic field} = \text{“electric field seen from a moving frame”}

Magnetism is not a separate force; it is the frame-dependent slice of FμνF_{\mu\nu}.


5.5.10 Quick recap

  • The field is FμνF_{\mu\nu}, the sources are JμJ^\mu, and the potential is AμA^\mu
  • Maxwell’s equations: μFμν=μ0Jν\partial_\mu F^{\mu\nu}=\mu_0 J^\nu and μF~μν=0\partial_\mu \tilde F^{\mu\nu}=0
  • Lorentz force: dpμ/dτ=qFμ νuνdp^\mu/d\tau = q F^\mu{}_{\ \nu} u^\nu
  • Boosts mix E\boldsymbol E and B\boldsymbol B, but the invariants I1,I2\mathcal I_1,\mathcal I_2 stay fixed
  • Energy–momentum flows with TμνT^{\mu\nu}; S=(1/μ0)E×B\boldsymbol S=(1/\mu_0)\boldsymbol E\times\boldsymbol B

In summary: Electromagnetism is not two fields but one tensor FμνF_{\mu\nu}. Lorentz transformations change how E\boldsymbol E and B\boldsymbol B look, not what the physics is. Writing everything in four-dimensional form makes covariance obvious, sharpens what’s invariant, and turns long vector calculus into compact geometry.


5.6 Relativistic Optics and Technology

Special relativity reshapes how light looks when sources or observers move fast. The result is a small toolkit—Doppler, aberration, and beaming—that explains phenomena from quasar jets to accelerator light and enables technologies from GPS to synchrotron X-ray sources. This section packages the main formulas, shows what is invariant, and sketches applications.


5.6.1 Relativistic Doppler: frequency shift as kinematics

For motion strictly along the line of sight with speed vv (define βv/c\beta \equiv v/c), the observed frequency is related to the emitted one by the Doppler factor

D1+β1β\mathcal D \equiv \sqrt{\frac{1+\beta}{1-\beta}}

Approaching sources are blueshifted, receding are redshifted:

νobs=Dνemit\nu_{\text{obs}}=\mathcal D\,\nu_{\text{emit}}

At small speeds this reduces to the classical shift. For purely transverse relative motion (source moves sideways), there is still a redshift due to time dilation alone:

νobs=νemitγ,γ=11β2\nu_{\text{obs}}=\frac{\nu_{\text{emit}}}{\gamma},\qquad \gamma=\frac{1}{\sqrt{1-\beta^2}}

5.6.2 Aberration of light: where the rays appear to come from

Motion squeezes apparent angles toward the direction of travel. If a light ray makes angle θ\theta with the xx-axis in SS and θ\theta' in SS' that moves at speed vv along +x+x, then

cosθ=cosθβ1βcosθ\cos\theta'=\frac{\cos\theta-\beta}{1-\beta\cos\theta}

Equivalently,

tan ⁣(θ2)=1β1+β  tan ⁣(θ2)\tan\!\left(\frac{\theta'}{2}\right)=\sqrt{\frac{1-\beta}{1+\beta}}\;\tan\!\left(\frac{\theta}{2}\right)

Starlight appears displaced (aberration), and radiation patterns are forward-beamed at high γ\gamma.


5.6.3 Beaming and intensity transformation

Specific intensity satisfies a powerful invariant:

Iνν3 is Lorentz invariant\frac{I_\nu}{\nu^3}\ \text{is Lorentz invariant}

Hence for a boost with Doppler factor D\mathcal D,

Iν=D3Iν,ν=DνI_\nu = \mathcal D^{3}\, I'_{\nu'},\qquad \nu=\mathcal D\,\nu'

Integrating over frequency (bolometric intensity) gives

I=D4II=\mathcal D^{4}\,I'

This headlight effect makes relativistic jets look dramatically brighter when pointed near our line of sight and dimmer when pointed away. Photons are not created; phase-space density is simply redistributed by the boost.


5.6.4 Synchrotron and curvature radiation

Charged particles forced into curved paths radiate. In circular motion of radius ρ\rho at ultra-relativistic speed, the spectrum peaks near the critical frequency

ωc=32γ3cρ\omega_c=\frac{3}{2}\,\gamma^{3}\,\frac{c}{\rho}

The total radiated power for motion with curvature ρ\rho scales strongly with γ\gamma,

Pq2cρ2γ4P \propto \frac{q^{2} c}{\rho^{2}}\,\gamma^{4}

(up to the standard SI prefactor). Two key consequences: (i) high-energy electrons in magnetic fields produce broad-band synchrotron radiation (radio to X-ray), and (ii) storage rings and undulators become bright, tunable light sources (synchrotron facilities, FELs) used for protein crystallography, materials science, and nanolithography.


5.6.5 Brightness temperatures and spectra (quick rules)

For a power-law emitter with IνναI'_{\nu'}\propto \nu'^{\,\alpha}, the observed spectrum under a boost obeys

IνD3+αναI_\nu \propto \mathcal D^{3+\alpha}\,\nu^{\,\alpha}

Isotropic rest-frame emission becomes anisotropic: apparent opening angles scale like 1/γ\sim 1/\gamma, and variability timescales shrink by 1/D1/\mathcal D (time compression). These scalings underlie inferences of bulk Lorentz factors in blazars and gamma-ray bursts.


5.6.6 Radar and lidar at speed: closing speed is not “just add”

When platforms move relativistically, measured beat frequencies in radar/lidar follow the same Doppler algebra. For monostatic radar with target closing speed vv along the line of sight, the round-trip shift accumulates twice:

νbeat(D21)νtx\nu_{\text{beat}} \approx \left(\mathcal D^{2}-1\right)\,\nu_{\text{tx}}

In the low-vv limit this reduces to the familiar 2v/c2v/c formula. At high vv, use the exact D\mathcal D to avoid systematic bias.


5.6.7 GPS and clocks: engineering with relativity

Modern navigation is a timing game. In special relativity, a clock moving at speed vv ticks slower by γ\gamma:

ΔtsatΔtground1γ1v22c2\frac{\Delta t_{\text{sat}}}{\Delta t_{\text{ground}}} \approx \frac{1}{\gamma} \approx 1-\frac{v^{2}}{2c^{2}}

Satellite clocks thus run slow due to orbital speed (special relativity) but fast due to higher gravitational potential (general relativity, §5.8). Flight software pre-biases clock frequencies and continuously corrects time using relativistic models; without these, position errors would grow to kilometers per day.


5.6.8 Imaging at relativistic speeds (thought experiments to lab)

If a camera moves near cc past a scene, aberration funnels most photons into a forward cone; Doppler shifts move spectra; and Terrell rotation (a geometry effect) makes objects appear rotated rather than squashed. Lab analogs show pieces of this: relativistic electron bunches “see” laser pulses Doppler-upshifted into X-rays (inverse Compton sources), enabling compact X-ray generation.


5.6.9 Cherenkov versus Doppler (don’t mix them)

Cherenkov radiation occurs when a charge moves through a medium faster than light’s phase speed in that medium. It is not superluminal in vacuum and does not violate relativity. In contrast, relativistic Doppler and aberration are vacuum kinematics. Both appear in detectors and astrophysics, but they answer different questions.


5.6.10 Practical crib sheet

  • Doppler factor: D=1+β1β\displaystyle \mathcal D=\sqrt{\frac{1+\beta}{1-\beta}}; use νobs=Dνemit\nu_{\text{obs}}=\mathcal D\,\nu_{\text{emit}}
  • Transverse Doppler: νobs=νemit/γ\displaystyle \nu_{\text{obs}}=\nu_{\text{emit}}/\gamma
  • Aberration: cosθ=cosθβ1βcosθ\displaystyle \cos\theta'=\frac{\cos\theta-\beta}{1-\beta\cos\theta}
  • Intensity: Iν/ν3I_\nu/\nu^{3} invariant  Iν=D3Iν\Rightarrow\ I_\nu=\mathcal D^{3} I'_{\nu'}; bolometric I=D4II=\mathcal D^{4} I'
  • Synchrotron: ωc=(3/2)γ3c/ρ\displaystyle \omega_c=(3/2)\gamma^{3} c/\rho; power γ4/ρ2\propto \gamma^{4}/\rho^{2}

In summary: Relativistic kinematics tilts light cones into everyday engineering. Doppler and aberration set what frequencies we detect and where sources appear; beaming and intensity invariants set how bright they look; synchrotron formulas govern radiation from bent charges. The same few relations explain cosmic jets, power ultra-bright light sources, and keep your phone’s map honest.


5.7 Equivalence Principle: Toward General Relativity

Special relativity rebuilt kinematics as geometry. Gravity finishes the renovation by changing the geometry itself. The doorway is Einstein’s equivalence principle: locally, a uniform gravitational field is indistinguishable from a uniformly accelerated frame. From this one idea flow gravitational redshift, the universality of free fall, and ultimately the field equations of general relativity.


5.7.1 Inertial vs gravitational mass: the clue

In Newtonian mechanics you meet two masses: inertial mim_{\mathrm i} (resistance to acceleration) and gravitational mgm_{\mathrm g} (how strongly gravity pulls). Experiments show they are equal to astonishing precision, so motion in a gravitational field is independent of composition. This universality of free fall hints that gravity is not a force in the usual sense but a property of spacetime affecting all matter the same way.

Einstein took the equality mg=mim_{\mathrm g}=m_{\mathrm i} as a postulate and asked: what worldview makes this identity natural?


5.7.2 Einstein’s elevator: redshift and light bending—without gravity

Imagine a small laboratory (an elevator) in deep space, initially at rest. Now fire its rockets so it accelerates upward with constant acceleration aa. A laser at the floor emits upward; by the time the light reaches the ceiling, the ceiling has picked up speed and sees a Doppler redshift. Over a height hh, the fractional shift is

Δννahc2\frac{\Delta \nu}{\nu} \approx -\,\frac{a h}{c^{2}}

Replace aa by gg and you have the gravitational redshift in a uniform field:

Δννghc2\frac{\Delta \nu}{\nu} \approx -\,\frac{g h}{c^{2}}

Same lab, different stunt: toss a horizontal light pulse across the cabin. During the flight time t/ct\approx \ell/c, the cabin accelerates downward relative to the beam, so the light appears to curve. By equivalence, light should also bend near a mass. The elevator thought experiment already whispers the core GR predictions.


5.7.3 Local inertial frames: gravity can be transformed away (but only locally)

At any spacetime event you can choose coordinates that make physics look like special relativity to first order. Technically: at a point PP there exist local inertial coordinates with

gμν(P)=ημν,αgμν(P)=0g_{\mu\nu}(P)=\eta_{\mu\nu},\qquad \partial_\alpha g_{\mu\nu}(P)=0

In such a frame, freely falling test bodies move on straight lines through PP and clocks behave as in SR—locally. You cannot make gravity vanish everywhere because second derivatives of the metric (curvature) remain. Those second derivatives encode tidal effects.


5.7.4 Geodesics: free fall as straightest possible paths

In GR, free fall is not “force-driven” motion; it is geodesic motion—extremal proper time curves in a curved manifold. The equation of motion is

d2xμdτ2+Γμαβdxαdτdxβdτ=0\frac{d^{2}x^{\mu}}{d\tau^{2}} + \Gamma^{\mu}{}_{\alpha\beta}\,\frac{dx^{\alpha}}{d\tau}\,\frac{dx^{\beta}}{d\tau} = 0

Here Γμαβ\Gamma^{\mu}{}_{\alpha\beta} are the Christoffel symbols built from gμνg_{\mu\nu}. In flat spacetime they can be set to zero globally; in curved spacetime they vanish only at a point by going to a local inertial frame.


5.7.5 Curvature and tides: geodesic deviation

Two nearby freely falling particles with separation ξμ\xi^\mu and 4-velocity uμu^\mu obey the geodesic deviation equation

D2ξμDτ2=Rμναβuνξαuβ\frac{D^{2}\xi^{\mu}}{D\tau^{2}} = -\,R^{\mu}{}_{\nu\alpha\beta}\,u^{\nu}\,\xi^{\alpha}\,u^{\beta}

RμναβR^{\mu}{}_{\nu\alpha\beta} is the Riemann curvature tensor. If R=0R=0, freely falling worldlines remain parallel (no tides). Nonzero curvature makes them focus or defocus—that is gravity in GR. The “gravitational field” is not a vector; it is geometry.


5.7.6 From equivalence to field equations

Einstein asked for the simplest field equations that

  1. reduce to Newtonian gravity when fields are weak and speeds are slow,
  2. conserve energy–momentum, and
  3. are generally covariant (valid in any coordinates).

The answer is the Einstein field equation

GμνRμν12Rgμν=8πGc4TμνG_{\mu\nu} \equiv R_{\mu\nu} - \tfrac{1}{2} R\,g_{\mu\nu} = \frac{8\pi G}{c^{4}}\,T_{\mu\nu}

Optionally one may add a cosmological constant Λ\Lambda on the left (+Λgμν)(+\Lambda g_{\mu\nu}); we omit it below for simplicity. TμνT_{\mu\nu} is the stress–energy tensor of matter and fields. Geometry GμνG_{\mu\nu} on the left, sources TμνT_{\mu\nu} on the right—the slogan is “matter tells spacetime how to curve; spacetime tells matter how to move.”


5.7.7 Newtonian limit: recovering Poisson’s law

Consider a static, weak field so that the metric component is

g001+2Φc2g_{00} \approx 1 + \frac{2\Phi}{c^{2}}

where Φ\Phi is the Newtonian potential with Φ/c21|\Phi|/c^{2}\ll 1. Insert this ansatz into the field equations and keep lowest order in 1/c21/c^{2}. One obtains

2Φ=4πGρ\nabla^{2}\Phi = 4\pi G\,\rho

and free-fall trajectories reduce to a=Φ\boldsymbol a = -\nabla\Phi. GR passes the sanity check: it contains Newton as the slow-motion, weak-field limit.


5.7.8 Gravitational time dilation and redshift (proper time in a static field)

For a stationary metric, a clock at fixed spatial coordinates ticks proper time

dτ=g00dtd\tau = \sqrt{g_{00}}\,dt

Near Earth, with g001+2Φ/c2g_{00}\approx 1+2\Phi/c^{2} and small potential differences ΔΦgΔh\Delta\Phi\approx g\,\Delta h, the fractional rate shift between altitudes is

ΔννΔΦc2gΔhc2\frac{\Delta \nu}{\nu} \approx -\,\frac{\Delta \Phi}{c^{2}} \approx -\,\frac{g\,\Delta h}{c^{2}}

This is why GPS satellite clocks (higher Φ\Phi) run faster than identical ground clocks and must be pre-biased and corrected.


5.7.9 Accelerated frames and Rindler coordinates (flat spacetime warm-up)

Uniformly accelerated observers in flat spacetime cover a Rindler wedge. One convenient chart in 1+11+1 dimensions uses parameters (ρ,η)(\rho,\eta) related to inertial (ct,x)(ct,x) by

ct=ρsinh ⁣(aηc),x=ρcosh ⁣(aηc)ct = \rho\,\sinh\!\left(\frac{a\eta}{c}\right),\qquad x = \rho\,\cosh\!\left(\frac{a\eta}{c}\right)

In these coordinates the metric reads

ds2=(aρc)2c2dη2dρ2ds^{2} = \left(\frac{a\rho}{c}\right)^{2} c^{2} d\eta^{2} - d\rho^{2}

Clocks at different ρ\rho tick at different rates, mimicking gravitational redshift. Equivalence tells us that many “accelerated frame” effects have gravitational analogs when curvature is small over the lab.


5.7.10 Light deflection, delay, and perihelion (previews)

Full GR predicts quantitative effects that can be phrased through the metric. For a light ray grazing a mass MM with impact parameter bb, the bending angle is

Δθ4GMc2b\Delta\theta \approx \frac{4 G M}{c^{2} b}

Signals skimming a mass experience an extra Shapiro time delay tied to g00g_{00} and spatial curvature. Bound orbits in curved spacetime precess; Mercury’s perihelion shift (after subtracting planetary perturbations) matches GR. We will collect classic tests in §5.9.


5.7.11 What the equivalence principle really asserts

  • Weak equivalence (universality of free fall): test bodies with negligible self-gravity follow the same trajectories given identical initial conditions in a given gravitational field
  • Einstein equivalence: in any local freely falling frame, non-gravitational physics reduces to special relativity; outcomes are independent of where and when the experiment is performed

Equivalence does not say gravity can be transformed away everywhere, nor that “gravity is a fictitious force.” It says that locally you can remove gravitational effects by free fall, leaving only curvature (tidal terms) as the invariant content.


5.7.12 Minimal problem kit

  • From a given metric gμνg_{\mu\nu}, compute dτ=g00dtd\tau=\sqrt{g_{00}}\,dt for clocks at rest and compare rates
  • For slow motion in a weak field, read off Φ\Phi from g001+2Φ/c2g_{00}\approx 1+2\Phi/c^{2} and recover Newtonian limits
  • To track free fall, set up the geodesic equation with Γμαβ(g)\Gamma^{\mu}{}_{\alpha\beta}(g) and integrate, or use conserved quantities from symmetries (Killing vectors)

In summary: The equivalence principle recasts gravity as geometry. Locally you can fall away the field, but you cannot eliminate curvature—tidal effects are the real, frame-invariant residue. Demanding a theory that honors equivalence, conserves energy–momentum, and recovers Newton leads to Einstein’s equation Gμν=(8πG/c4)TμνG_{\mu\nu}=(8\pi G/c^{4})T_{\mu\nu}. From redshift to light bending to clock rates in orbit, the elevator thought experiment graduates into a full-blown theory of curved spacetime.


5.8 Tests and Frontiers

Relativity is not a vibe check; it is an experiment magnet. From Mercury’s orbit to gravitational waves and black-hole images, the theory keeps passing increasingly savage tests. This section hits the classics, then pushes to frontiers where precision, strong gravity, or cosmic scales stress the model.


5.8.1 Classic tests in the weak field

Perihelion advance of Mercury. Newtonian gravity plus planetary perturbations misses a small excess. For a bound orbit with semi-major axis aa and eccentricity ee around mass MM, GR predicts per orbit

Δφ6πGMc2a(1e2)\Delta\varphi \approx \frac{6\pi G M}{c^{2} a\,(1-e^{2})}

Deflection of light. A light ray grazing a mass MM with impact parameter bb is bent by

Δθ4GMc2b\Delta\theta \approx \frac{4 G M}{c^{2} b}

This doubles the naive Newtonian-particle estimate because spacetime curvature bends both time and space.

Gravitational redshift. In a static weak field, clocks at higher potential tick faster. For small height difference hh near Earth with acceleration gg,

Δννghc2\frac{\Delta\nu}{\nu} \approx -\,\frac{g\,h}{c^{2}}

Shapiro time delay. Radar signals skimming a massive body arrive late compared with Euclidean expectations. For a round trip past mass MM with appropriate geometry,

ΔtGMc3ln ⁣(path distancesimpact parameter)\Delta t \propto \frac{G M}{c^{3}}\,\ln\!\left(\frac{\text{path distances}}{\text{impact parameter}}\right)

The logarithm is the telltale GR signature.


5.8.2 Equivalence principle tests

If all uncharged test bodies fall the same way, the Eötvös parameter must be tiny

η2a1a2a1+a2\eta \equiv \frac{2\,|a_{1}-a_{2}|}{a_{1}+a_{2}}

Modern torsion balances, atom interferometers, and lunar laser ranging constrain η\eta to be extremely small, keeping Einstein’s elevator happily indistinguishable from a lab in free fall.


5.8.3 Binary pulsars and gravitational radiation

A compact binary loses orbital energy to gravitational waves, shrinking its period. In GR the period derivative depends on the chirp mass McM_{c}

Mc(m1m2)3/5(m1+m2)1/5M_{c} \equiv \frac{(m_{1} m_{2})^{3/5}}{(m_{1}+m_{2})^{1/5}}

For a quasi-circular binary with orbital period PP,

P˙=192π5(GMcc3)5/3(2πP)5/3\dot P = -\,\frac{192\pi}{5}\left(\frac{G M_{c}}{c^{3}}\right)^{5/3}\left(\frac{2\pi}{P}\right)^{5/3}

Radio timing of pulsars matches this prediction and measures McM_{c} cleanly. Strong-field, radiative GR survives the check.


5.8.4 Direct gravitational waves and chirps

Interferometers measure a dimensionless strain h=ΔL/Lh=\Delta L/L from passing ripples. For a quasi-circular inspiral, the gravitational-wave frequency ff accelerates according to

f˙=965π8/3(GMcc3)5/3f11/3\dot f = \frac{96}{5}\,\pi^{8/3}\left(\frac{G M_{c}}{c^{3}}\right)^{5/3} f^{11/3}

From f(t)f(t) one extracts McM_{c} and distances without a cosmic distance ladder. Multi-detector triangulation localizes sources; polarization and phasing test GR’s waveform templates.


5.8.5 Black holes and strong gravity

Define the Schwarzschild radius

rs2GMc2r_{s} \equiv \frac{2 G M}{c^{2}}

Light rings, innermost stable circular orbits, and tidal forces near rsr_{s} probe the metric where curvature is large. Very long baseline interferometry resolves horizon-scale structure; stellar orbits around galactic nuclei map spacetime and test the no-hair paradigm through precessions and redshifts.


5.8.6 Frame dragging and geodetic effects

A spinning mass drags inertial frames. Around a body of angular momentum J\boldsymbol J, the gravitomagnetic precession rate of a gyroscope is roughly

ΩLTGJc2r3\boldsymbol{\Omega}_{\text{LT}} \sim \frac{G\,\boldsymbol J}{c^{2} r^{3}}

Geodetic precession arises from motion through curved spacetime. Satellite gyroscopes and tracking of satellite nodes measure both effects in Earth’s field, matching GR to good accuracy.


Time dilation and gravitational redshift are not optional in engineering. Satellite clock rates are pre-biased and corrected using SR and GR so that ranging equations remain consistent at meter levels and below. Fiber and microwave time-transfer links carry these relativistic corrections; atomic clock transport experiments confirm the predicted offsets.


5.8.8 Parameterized post-Newtonian (PPN) language

To compare alternatives to GR, weak-field observables are written with parameters like γ\gamma and β\beta. For example, light deflection is proportional to 1+γ1+\gamma; perihelion advance depends on a combination of γ\gamma and β\beta. GR predicts γ=β=1\gamma=\beta=1. Solar-system and pulsar tests pin these parameters extremely close to their GR values.


5.8.9 Propagation speed, dispersion, and graviton mass

If gravity propagated at a speed different from cc, or if the graviton had mass mgm_{g}, waves would disperse. A massive graviton obeys

E2=p2c2+mg2c4E^{2} = p^{2} c^{2} + m_{g}^{2} c^{4}

which makes higher-frequency waves arrive earlier than lower-frequency ones from the same source. Observed waveforms and multi-messenger events bound any dispersion very tightly, consistent with propagation at cc within tiny fractions.


5.8.10 Cosmological frontiers

On the largest scales, GR plus simple matter components yields expanding-universe solutions. Introducing a cosmological constant Λ\Lambda gives accelerated expansion. Geometry meets energy content via the Friedmann equation for scale factor a(t)a(t)

(a˙a)2=8πG3ρkc2a2+Λc23\left(\frac{\dot a}{a}\right)^{2} = \frac{8\pi G}{3}\,\rho - \frac{k c^{2}}{a^{2}} + \frac{\Lambda c^{2}}{3}

Distance–redshift relations, lensing, and structure growth test GR across cosmic time. Deviations would show up as tension between these probes.


5.8.11 Where new physics could appear

  • Strong-field bumps. Ringdown overtones from black-hole mergers test the spectrum expected from a Kerr metric
  • Equivalence principle edges. Ultralight fields or dark-sector couplings could induce composition-dependent accelerations or time variation in constants
  • Short-distance gravity. Tabletop experiments hunt for deviations from the inverse-square law at millimeter to micron scales
  • Lorentz symmetry. High-energy cosmic rays and precise timing look for tiny violations in the gravitational sector

So far, null results keep GR in front, but the parameter space keeps shrinking.


5.8.12 Minimal problem kit

  • Compute the Mercury-like precession with Δφ\Delta\varphi above and compare to a Newtonian ellipse over many orbits
  • From a grazing light path, estimate deflection and Shapiro delay using the formulas here
  • For a binary with given masses and period, use P˙\dot P and f˙\dot f to predict waveform evolution and inspiral time
  • Given two clock heights Δh\Delta h, evaluate the gravitational redshift and relate it to GPS bias settings

In summary: Relativity has survived a century of cross-examination: orbits precess, light bends and delays, clocks shift, binaries chirp, and horizons cast shadows exactly as the equations say. The frontiers now are precision and extremes—microns in the lab, megaparsecs in the cosmos, and curvatures near rsr_{s}. Wherever we push, the playbook is the same: compute invariants, respect the light cone, and let the metric tell matter how to move.