All Possible Paths: Sum-Over-Histories and Quantum Amplitudes
Look, if you want to understand quantum mechanics—really understand it, not just calculate with it—you need to forget everything you learned about particles taking definite paths. That’s classical thinking. It served Newton well for billiard balls and planets, but nature doesn’t work that way at the quantum level.
Here’s what I discovered when I was working on my PhD thesis at Princeton in the early 1940s: a particle going from point A to point B doesn’t take just one path. It takes every possible path simultaneously. The straight line. The curved arc. The detour to Jupiter and back. The absurd zigzag that violates every intuition you have about how things should move. All of them. At once.
This isn’t philosophy. This is calculation. And it works.
The Classical Action Principle
Let’s start with what the old guys knew. In classical mechanics, you can reformulate Newton’s laws using something called the action principle. Instead of forces and accelerations, you describe motion using a quantity called the action S, which is the integral of the Lagrangian L over time: S = ∫L dt, where L equals kinetic energy minus potential energy.
The beautiful thing about this formulation—and Euler, Lagrange, and Hamilton all recognized this—is that you can derive the actual path a particle takes by demanding that the action be stationary. That’s the principle of least action, though “stationary” is more accurate than “least.” The path nature chooses is the one where small variations don’t change the action to first order: δS = 0.
Think about throwing a ball from point A to point B. Classical mechanics says it follows a parabola—the trajectory that makes the action stationary. Any other path, whether it’s a straight line or some crazy loop-de-loop, has a different action. And classical mechanics says: nature doesn’t take those paths. Period.
But here’s what I realized: quantum mechanics doesn’t forbid those paths. It includes them all.
Interference of Paths
My key insight was this: instead of one path with δS = 0, consider all paths from A to B. Each path contributes an amplitude—a complex number—given by e^(iS/ℏ) where S is the action for that specific path and ℏ is Planck’s constant divided by 2π.
The total quantum amplitude to go from A to B is the sum—really an integral—over all possible paths:
ψ(B) = ∫ e^(iS[path]/ℏ) 𝒟[path]
This is the path integral. That 𝒟[path] notation means you’re integrating over an infinite-dimensional space of all conceivable paths. It’s not just summing over a few alternatives. It’s including the path that goes backward in time, the path that spirals around the moon, every crazy trajectory you can imagine.
Now here’s the crucial question: if every path contributes, why don’t we see quantum effects in everyday life? Why does a thrown baseball follow a parabola and not some quantum superposition of wild trajectories?
The answer is interference.
Look at the phase factor: e^(iS/ℏ). For macroscopic objects, the action S is enormous compared to ℏ. A baseball has an action on the order of joule-seconds, while ℏ is about 10^-34 joule-seconds. That means neighboring paths have phases that differ by billions and billions of radians. They oscillate wildly—some positive, some negative—and they cancel each other out through destructive interference.
Except near the classical path.
Near the path where δS = 0, something special happens. Neighboring paths have nearly the same action—they’re at a stationary point of S—so their phases align. You get constructive interference. This is the stationary phase approximation, and it’s why classical mechanics emerges from quantum mechanics in the appropriate limit.
When ℏ is small compared to S, the only paths that don’t cancel are those near the classical trajectory. But when S is comparable to ℏ—which happens for electrons, photons, and other quantum particles—many paths interfere constructively. You get tunneling through barriers, diffraction around obstacles, the double-slit interference pattern. All the quantum weirdness.
Take the double-slit experiment. An electron travels from the source to a point on the screen. Classically, it would go through one slit or the other. But quantum mechanically, it takes all paths—including paths through both slits. The paths through the left slit have one set of phases, paths through the right slit have another. When you add them up, you get interference: bright fringes where the phases align (Δφ = 2πn), dark fringes where they cancel (Δφ = π(2n+1)).
My path integral formulation makes this automatic. No mysterious wave-particle duality. No weird collapse of the wavefunction. Just sum over all paths, calculate the interference, and out comes the probability.
And yes, this is mathematically equivalent to Schrödinger’s equation. I proved that. But the physical picture is different—and for me, much clearer.
Feynman Diagrams and QED
The path integral gets even more interesting when you apply it to quantum field theory. Now you’re not just summing over paths of particles—you’re summing over all possible field configurations, including those where particles appear and disappear.
In quantum electrodynamics, electrons and photons are excitations of underlying fields. An electron can emit a photon, absorb one, or interact through virtual photon exchange. And you have to sum over all possible ways these interactions can happen.
That’s where my diagrams come in. Back in 1948, I developed a visual way to represent these quantum field theory calculations. Each diagram depicts one possible scenario: particles enter (external lines), interact (vertices), exchange virtual particles (internal lines), and exit. Lines have arrows: forward for electrons, backward for positrons. Wavy lines represent photons.
Every diagram corresponds to a mathematical amplitude. The rules are precise: each vertex contributes a factor proportional to the fine structure constant α ≈ 1/137. Each internal line gets a propagator. You integrate over all possible momenta for the virtual particles.
Then you sum all the diagrams to a given order in perturbation theory.
The simplest process—say, electron-electron scattering—has a leading diagram with one photon exchanged. Two vertices, amplitude proportional to α^2. But there are corrections: diagrams with two photons exchanged, or one photon with a virtual electron-positron loop. These have four vertices, so they’re down by a factor of α^4 compared to the leading term—about a 1% correction.
This is why QED works so well. The coupling is weak (α ≈ 1/137 is small), so the perturbation series converges rapidly. The first few diagrams give you incredible precision.
And the predictions are staggering. The electron’s magnetic moment—its g-factor—is predicted by QED to be 2(1 + α/(2π) + …) = 2.00231930436… We’ve calculated this to five loop diagrams. Experimentally, it’s measured to match that prediction to ten decimal places. Ten. That’s the most precisely tested theory in the history of science.
The Lamb shift, vacuum polarization, quantum corrections to Coulomb’s law—QED explains them all. And it’s all built on the path integral: sum over all field configurations, each weighted by e^(iS/ℏ), interfere them, and extract the amplitude.
Sure, there are technical issues. Loop diagrams give infinities, which we handle through renormalization—absorbing the divergences into redefinitions of mass and charge. It’s mathematically rigorous but philosophically unsettling. I never fully liked it, even though I got the Nobel Prize for this work in 1965.
But pragmatically? It works. Shut up and calculate, as I like to say.
Quantum Computing and Modern Applications
Now here’s where the path integral becomes more than just a theoretical tool. It’s the conceptual foundation for quantum computing.
Think about a quantum algorithm like Grover’s search or Shor’s factoring. The key insight is interference engineering. You set up a quantum superposition of all possible computational paths—all possible combinations of qubit states—and let them evolve. Some paths lead to the right answer, others don’t. The art is designing the evolution so that paths leading to wrong answers interfere destructively and cancel, while paths leading to the correct answer interfere constructively and amplify.
It’s exactly the same principle as the double-slit experiment or the path integral. You’re not computing along one trajectory. You’re computing along all trajectories in parallel, exploiting interference to pick out the answer.
Grover’s algorithm searches an unsorted database of N items in O(√N) time instead of the classical O(N). Shor’s algorithm factors large numbers in polynomial time instead of exponential. Both work because quantum mechanics explores all possibilities simultaneously and uses interference to filter the solution.
In a sense, this is what quantum mechanics has always been doing. When an electron finds the ground state of an atom, it’s not trying one orbital and then another. It’s sampling all configurations, and the interference structure of the path integral picks out the state with lowest energy.
Beyond computing, path integral methods underlie modern quantum simulations—something I proposed back in 1982. Classical computers struggle with quantum many-body systems because the Hilbert space grows exponentially. But a quantum computer represents quantum states naturally. Simulating strongly-correlated electrons, quantum chemistry, high-temperature superconductors—these become tractable.
There’s also path integral Monte Carlo, a numerical technique for finite-temperature quantum systems. You can’t sum over infinitely many paths exactly, but you can sample them stochastically, estimate the partition function, and calculate thermodynamic properties. It’s used for liquid helium, ultracold atoms, quantum phase transitions.
The path integral unifies all of this. It’s the same formalism whether you’re describing an electron in a potential well, a quantum field theory like QED, or a quantum computer running Shor’s algorithm. Particles and fields, non-relativistic and relativistic, single-particle and many-body. Sum over histories. Interfere the amplitudes. Calculate.
Why It Matters
The path integral formulation does more than reproduce the results of wave mechanics. It reveals something profound about quantum reality: nature doesn’t choose one history. It explores all histories. The appearance of a definite outcome emerges from the interference pattern of infinitely many possibilities.
Philosophically, this dissolves determinism in a subtle way. It’s not that the future is random—quantum mechanics is perfectly deterministic in how the wavefunction evolves. It’s that the wavefunction itself is a superposition of every conceivable path, and what we call “measurement” is the collapse of that superposition into a definite result.
Some people—Everett, DeWitt—took this to its logical conclusion: maybe all paths are real, and measurement just splits the universe into branches. The many-worlds interpretation. I’m agnostic. Interpreting quantum mechanics is above my pay grade. I care about calculation and prediction.
But I will say this: the path integral teaches you humility. When a particle goes from A to B, you might think it took the obvious route. Quantum mechanics says it took every route. The straight line and the detour to Andromeda. The sensible and the absurd. And only through the conspiracy of interference—the way those complex amplitudes add up—do you get the probabilities you measure.
That’s beautiful. That’s strange. And that’s nature.
Source Notes
9 notes from 2 channels
Source Notes
9 notes from 2 channels