Feynman’s Path Integral Approach to Quantum Mechanics

PRMJ
Technological Singularity
10 min readApr 14, 2024
Richard Feynman

Quantum mechanics, a fundamental theory in physics, describes the behavior of particles at the smallest scales. While traditional formulations like the Schrödinger equation have offered deep insights, Richard Feynman introduced a different and highly influential method in the 1940s known as the path integral approach. This method has since reshaped our understanding of particle behavior, providing a new perspective that complements other quantum theories.

Feynman’s approach is fascinating because it proposes that particles don’t travel from one point to another along a single path; instead, they explore every possible path simultaneously. In classical mechanics, a particle has a definite path, determined by which path minimizes the action — a concept that measures the dynamics of the system. However, in the quantum world, things aren’t so straightforward. According to Feynman, every conceivable path contributes to the probability of where a particle will be found.

To calculate these probabilities, Feynman assigned a number, called a probability amplitude, to each path. This number isn’t just any ordinary number but involves complex numbers and exponentials. Specifically, the probability amplitude for each path is proportional to the exponential of the action of that path, multiplied by a complex phase. This might sound complicated, but the core idea is that the action of a path influences how much that path contributes to the final destination of the particle.

According to Feynman’s theory a photon doesn’t follow just one route. Instead, it takes every possible route simultaneously. Some paths are straight, and some are more indirect than others. The probability of finding the photon at a particular spot is determined by adding up the complex amplitudes of all these paths. Paths with similar actions will add constructively and increase the likelihood of the photon arriving along those paths, while others will cancel out due to differing phases, demonstrating quantum interference.

This sum over all possible histories is not a straightforward addition, as it involves integrating over an infinite number of possible paths, a process made manageable through advanced mathematical techniques. The path integral approach excels in systems where actions are dominated by quadratic terms, making it solvable and useful for exact solutions and approximations alike.

Let’s consider a practical example that illustrates Feynman’s path integral approach in a quantum mechanical context, using the analogy of light traveling from a distant star to Earth — a scenario that might seem straightforward at first but is intriguing when viewed through the lens of quantum mechanics.

Light’s Journey from a Star to Earth

Imagine light emitted from a star many light years away. In classical physics, we would predict that light travels in a straight line from the star to Earth, following the path of least action, a direct route that minimizes energy used over time. This is a simple and efficient path, much like choosing the shortest walking route in a city map.

However, according to Feynman’s path integral formulation, the photon emitted by the star does not take just one path. Instead, it takes every conceivable path to reach Earth. Some of these paths are direct, some zigzag wildly through space, and others may loop around other celestial bodies before heading toward Earth.

Each of these paths contributes to the overall probability amplitude of the photon’s arrival at a specific location on Earth. The paths closer to the classical straight line, being simpler and requiring less action (in terms of interactions and deviations), contribute more significantly to the probability amplitude. These are analogous to major roads on a map that are typically faster and more direct. Conversely, the more convoluted paths, which might involve bouncing off interstellar objects or taking long detours, contribute much less to the final probability amplitude, much like a convoluted detour on a city map that is less likely to be chosen for a quick commute.

To visualize these multiple paths, one could use Feynman diagrams, which would depict not only the direct interactions but also possible interactions with other particles along the way. These diagrams simplify the complex probabilities into manageable visualizations, where each vertex or interaction point on the diagram represents a potential change in the path or state of the photon.

Ultimately, the photon’s behavior is a superposition of all these paths, and the path that we observe (the photon arriving on Earth from the star) is only one of many possibilities, influenced by the sum of the amplitudes of all possible paths. This emphasizes the quantum nature of reality, where every possible history has its contribution, and our classical intuition about the simplest path is only part of a much richer, probabilistic picture.

This example encapsulates the essence of Feynman’s path integral approach and demonstrates its radical departure from classical mechanics, illustrating a universe that is fundamentally probabilistic rather than deterministic.

The Double Slit Experiment

The double slit experiment is a classic example of Feynman’s path integral formulation in action and beautifully demonstrates the core principles of quantum mechanics, particularly quantum superposition and interference.

In the double slit experiment, particles (such as electrons, photons, or even larger molecules) are fired towards a barrier that has two slits. After passing through the slits, the particles hit a detection screen behind the barrier. If particles behaved like classical particles, you would expect to see two distinct bands directly opposite the two slits on the detection screen, corresponding to the paths through each slit.

However, what is actually observed is a pattern of many bands or fringes of varying brightness. This pattern results from interference, indicative of the particles exhibiting wave-like properties. According to quantum mechanics, each particle passes through both slits simultaneously, not as a single localized entity but as a wave spreading out and passing through both slits.

Feynman’s path integral approach provides a profound explanation for this phenomenon. It suggests that each particle explores all possible paths from the source to the detection screen. The paths include going through the first slit, going through the second slit, possibly bouncing off the edges of the slits, and every other imaginable path. The interference pattern arises because the probability amplitudes associated with these paths combine.

Some paths will be in phase and constructively interfere (enhancing the probability of the particle arriving at specific points on the screen), while others will be out of phase and destructively interfere (reducing the probability of the particle arriving at other points). The bright fringes on the screen correspond to points of constructive interference (where paths add up in phase), and the dark areas correspond to destructive interference (where paths cancel out).

This experiment highlights the radical departure from classical intuition: particles in quantum mechanics do not have well-defined positions or paths before measurement. Instead, their behavior is better described by a sum over all possible paths, with the final observed outcome being a result of the interference of all these paths’ probability amplitudes. This experiment directly supports the idea that quantum entities like photons and electrons behave as both particles and waves, depending on how they are observed, a central concept in quantum mechanics and particularly illustrative of Feynman’s contributions.

Beyond providing a new way to calculate quantum phenomena, the path integral formulation has found essential applications in various physics fields. In quantum field theory, it helps calculate how particles scatter and interact. In statistical mechanics, it offers methods to analyze systems at thermal equilibrium. Moreover, the approach has potential implications for quantum gravity, where non-perturbative methods are necessary.

Feynman’s path integral approach has dramatically expanded our toolkit for understanding and calculating quantum effects. By conceptualizing particle motion as a sum over all conceivable paths, Feynman not only offered a new perspective on quantum mechanics but also paved the way for advancements in several branches of theoretical physics. His ideas continue to influence how we understand the fundamental nature of the universe, demonstrating the power of looking at old problems through new lenses.

Fundamental Example

A problem that demonstrates the principles of Feynman’s path integral formulation in a basic setting is the calculation of the probability amplitude for a particle to move from one point to another in a given time. Let’s consider a simple scenario: a free particle moving from point A to point B without any potential field or interactions.

Problem Setup:

  • Initial Condition: A particle starts at
  • Final Condition: The particle ends at

Objective: Calculate the probability amplitude for the particle to move from A​ to B under the path integral formulation.

Define the Action: The action S for a free particle (no potential energy) is given by:

where x dot is the velocity of the particle, and m is its mass.

Path Integral: The probability amplitude K for the particle to transition from A to B​ is given by the path integral

where Dₓ indicates integration over all possible paths x(t) that start at A​ and ends at B​, and ℏ is the reduced Planck constant.

Evaluating the Path Integral: For a free particle, the path integral can be solved exactly. The paths that contribute significantly are those close to the classical path that minimizes the action S. In this simple case, the classical path is a straight line motion from A to B​ with constant velocity

The action along this path is:

Solution: The exact solution for the probability amplitude in this case is given by

The probability amplitude K represents the quantum mechanical amplitude for the particle to travel from A​ to B​ in the time interval​, considering all possible paths but with dominant contributions from the classical path. This example showcases the basic application of Feynman’s path integral in calculating quantum amplitudes for simple quantum systems.

In the path integral formulation of quantum mechanics, particularly for simple systems like a free particle (where there are no forces acting on the particle), the most likely path — the path that contributes most significantly to the path integral — is a straight line path. This path corresponds to the classical path that the particle would follow according to classical mechanics.

In classical mechanics, the path that a system takes between two points is the one that minimizes the action (Principle of Least Action), a quantity which is defined as the integral over time of the Lagrangian of the system. The Lagrangian for a free particle is

The action for any path from point A to point B over a time t is given by:

The path that minimizes this action under the condition of fixed endpoints is a straight line with constant velocity.

In the quantum realm, even though all paths are considered, the path integral formulation suggests that the sum over all possible paths between two points is dominated by those paths near the one that minimizes the action — the classical path. This is because contributions from paths far from the classical path tend to interfere destructively and cancel out, reducing their impact on the overall probability amplitude.

The probability amplitude for a path is weighted by the exponential of the action (multiplied by i/ℏ),

Near the classical path, where the action is minimal, these contributions add constructively.

Although quantum mechanics allows for the possibility of a particle taking bizarre and non-intuitive paths (like looping or detouring significantly from the direct route), the overwhelming contributions come from paths that are close to what we expect in classical mechanics when no forces are present.

The straight-line path not only makes physical sense (being the shortest and requiring the least change in kinetic energy) but also aligns with the predictions of classical physics in the limit of large quantum numbers or large masses (correspondence principle).

This agreement between the classical path and the dominant contribution in the path integral is a beautiful aspect of quantum mechanics that shows how classical mechanics emerges from quantum principles under appropriate conditions. It illustrates how Feynman’s path integral approach bridges the gap between quantum behavior and classical laws.

General Relativity and Quantum Gravity

In the context of quantum gravity, the focus shifts from traditional forces to the dynamics and curvature of spacetime itself, which is a significant departure from how other forces (like electromagnetism, weak, and strong nuclear forces) are described in quantum field theory.

In Einstein’s theory of general relativity, gravity is not treated as a force between masses, as it was in Newtonian physics. Instead, it is described as the effect of the curvature of spacetime caused by mass and energy. Objects move along paths called geodesics, which are essentially the "straightest" possible paths in this curved spacetime. The presence of mass or energy bends spacetime, and this curvature dictates how objects move.

Quantum gravity seeks to extend this concept into the quantum realm, where quantum effects become significant, such as at extremely small scales close to the Planck length (10^-35 meters) or in conditions of immense gravitational fields like those near black holes or during the early moments of the Big Bang.

The extension of path integrals to quantum gravity represents one of the more ambitious and complex areas in theoretical physics. Path integrals, which were initially developed for quantum mechanics and quantum field theory, offer a way to calculate the probability amplitudes by summing over all possible histories of a system. When applied to gravity, this approach attempts to integrate over all possible geometries of spacetime itself, a concept that fundamentally challenges and extends our understanding of both quantum mechanics and general relativity.

Reference

“The Quantum Entanglement. (2020, April). Feynman’s Path Integral Formulation and Its Implications. Retrieved from https://thequantumentanglement.blogspot.com/2020/04/feynmans-path-integral-formulation-and.html"

“Feynman’s Path Integral Explained with Basic Calculus” by Swapnonil Banerjee, Ph.D.: Banerjee, S. (2023). Feynman’s Path Integral Explained with Basic Calculus. SwaNi.

--

--