The Unseen Architect: How the Hamiltonian Equation Forges the Quantum World
In the heart of the perplexing and beautiful realm of quantum mechanics lies a single, powerful entity: the Hamiltonian. It is an unassuming mathematical operator, yet it holds the key to unlocking the very essence of quantum systems. From the intricate dance of electrons in an atom to the collective behavior of materials and the revolutionary promise of quantum computers, the Hamiltonian stands as the central architect, dictating the evolution and properties of the quantum world. This profound equation, a legacy of classical physics reimagined for a new, probabilistic reality, is more than just a tool for calculation; it is a window into the fundamental nature of energy, time, and the very fabric of existence.
From Celestial Mechanics to the Quantum Leap: The Genesis of the Hamiltonian
The story of the Hamiltonian begins not in the strange world of the infinitesimally small, but in the elegant and deterministic realm of classical mechanics. In the 1830s, the Irish mathematician and physicist Sir William Rowan Hamilton sought a more profound and unified way to describe the motion of objects, from falling apples to orbiting planets. At the time, Newtonian mechanics, with its focus on forces and accelerations, reigned supreme. However, Hamilton envisioned a description based on a more fundamental quantity: energy.
Hamilton's stroke of genius was to formulate a new framework for mechanics, now known as Hamiltonian mechanics, which described a system's state not by its position and velocity, but by its position and momentum. He introduced a function, which he called the "characteristic function," but which is now universally known as the Hamiltonian, that represents the total energy of a system—the sum of its kinetic and potential energies. This seemingly simple shift in perspective had profound consequences. Hamilton's equations of motion, derived from this Hamiltonian function, provided a new and often more powerful way to predict the evolution of a physical system.
One of the most elegant aspects of Hamilton's formulation was its use of a concept called "phase space." For a single particle moving in one dimension, phase space is a two-dimensional plane where the particle's position is plotted on one axis and its momentum on the other. The state of the system at any given time is represented by a single point in this space, and as the system evolves, this point traces out a path. Hamilton's equations govern how this point moves through phase space, providing a complete and deterministic picture of the system's dynamics.
A crucial feature of Hamiltonian mechanics is the symmetric treatment of position and momentum, known as "conjugate variables." This mathematical elegance would prove to be a crucial stepping stone for the development of quantum mechanics.
Another key aspect of Hamilton's work was his development of the "Hamilton-Jacobi equation," a powerful tool for solving problems in classical mechanics. This equation, and the broader framework of Hamiltonian mechanics, revealed a deep connection between mechanics and optics. Hamilton's "optico-mechanical analogy" drew a parallel between the paths of particles in a potential field and the paths of light rays in a medium with a varying refractive index. This analogy, which connected the principles of Maupertuis (for mechanics) and Fermat (for optics), suggested that both particles and light rays follow paths that minimize a certain quantity. This idea of an underlying wave-like nature to mechanics was a prescient hint of the revolution to come.
As the 20th century dawned, the classical world of Newton and Hamilton began to crumble at its foundations. The discovery of phenomena like black-body radiation, the photoelectric effect, and atomic spectra revealed a world that was anything but continuous and deterministic. This was the quantum world, and it demanded a new physics.
In the 1920s, a new generation of physicists, including Werner Heisenberg, Erwin Schrödinger, and Paul Dirac, laid the foundations of quantum mechanics. Schrödinger, in particular, was deeply influenced by Hamilton's work. He recognized that Hamilton's optico-mechanical analogy provided a natural starting point for a new "wave mechanics." If light could behave as both a wave and a particle, perhaps the same was true for matter.
Schrödinger's great insight was to take the classical Hamiltonian and transform it into a quantum mechanical operator. In classical mechanics, the Hamiltonian is a function of position and momentum, which are simply numbers. In quantum mechanics, position and momentum are no longer simple numbers but are instead represented by mathematical operators that act on a "wave function" (Ψ). The wave function, in Schrödinger's theory, contains all the information about a quantum system.
The process of "quantization" involved replacing the classical position and momentum variables in the Hamiltonian with their corresponding quantum operators. For a single particle moving in one dimension, the classical Hamiltonian is:
H = p²/2m + V(x)where p is the momentum, m is the mass, and V(x) is the potential energy.
In quantum mechanics, this becomes the Hamiltonian operator, often denoted with a "hat" to distinguish it from its classical counterpart:
Ĥ = p̂²/2m + V̂(x̂)Here, p̂ and x̂ are the momentum and position operators, respectively.
This quantum Hamiltonian is the heart of the Schrödinger equation, one of the most fundamental equations in all of physics. The time-independent Schrödinger equation, in particular, takes the form of an eigenvalue equation:
Ĥψ = Eψwhere Ĥ is the Hamiltonian operator, ψ is the wave function (or "eigenstate"), and E is a number representing the total energy of the system (the "eigenvalue").
This equation is a declaration of profound significance: when the Hamiltonian operator acts on a particular wave function, it returns the same wave function multiplied by a constant, which is the energy of that state. The possible energy values, the "energy spectrum" of the system, are the set of all possible eigenvalues. These are the only allowed energies that the quantum system can possess, a concept known as energy quantization.
The transition from classical to quantum mechanics was also elegantly captured by Paul Dirac, who showed that the Poisson brackets of classical Hamiltonian mechanics have a direct quantum analogue in the commutators of quantum operators. In classical mechanics, the time evolution of any quantity A is given by its Poisson bracket with the Hamiltonian. In quantum mechanics, the time evolution of an operator is given by its commutator with the Hamiltonian. This deep connection, where {A, H} in the classical world becomes (1/iħ)[Â, Ĥ] in the quantum world, solidified the Hamiltonian's central role in the new theory.
Thus, the Hamiltonian, born from the desire for a more elegant description of classical motion, found its true and ultimate purpose as the master key to the quantum realm. It provided the mathematical framework that allowed physicists to bridge the gap between the old, deterministic world and the new, probabilistic one.
The Hamiltonian in Action: Solving for Quantum Reality
The abstract beauty of the Hamiltonian operator comes to life when it is applied to specific quantum systems. By constructing the Hamiltonian for a given situation—defining its kinetic and potential energy terms—and then solving the Schrödinger equation, we can predict the system's allowed energies and the probability of finding a particle in a particular location. Let's explore some of the most fundamental and illustrative examples.
The Particle in a Box: A Quantum Prison
Imagine a particle confined to a one-dimensional "box" of length L, with infinitely high walls. Inside the box, the potential energy is zero, but at the walls and beyond, it is infinite, meaning the particle can never escape. This is one of the simplest, yet most instructive, problems in quantum mechanics.
The Hamiltonian for this system is straightforward. Since the potential energy V(x) is zero inside the box, the Hamiltonian consists only of the kinetic energy term:
Ĥ = - (ħ²/2m) (d²/dx²)where ħ is the reduced Planck constant and m is the mass of the particle.
Solving the time-independent Schrödinger equation (Ĥψ = Eψ) with this Hamiltonian, and applying the "boundary conditions" that the wave function must be zero at the walls of the box (since the particle cannot be outside the box), leads to a remarkable result. Only certain wave functions, and therefore only certain energies, are allowed.
The allowed wave functions, or eigenstates, are standing waves, much like the vibrations of a guitar string fixed at both ends:
ψn(x) = √(2/L) sin(nπx/L)
where n is a positive integer (1, 2, 3, ...), known as the quantum number.
Each of these eigenstates has a corresponding energy eigenvalue:
En = (n²h²)/(8mL²)where h is Planck's constant.
These results reveal some of the most fundamental features of quantum mechanics:
- Quantization of Energy: The energy of the particle is not continuous but is restricted to discrete, quantized levels. The particle can have energy E₁, E₂, E₃, and so on, but never an energy value in between. This is a direct consequence of the wave-like nature of the particle and the confinement imposed by the box.
- Zero-Point Energy: The lowest possible energy, the "ground state" energy (E₁), is not zero. This is a stark contrast to classical mechanics, where a particle can be at rest with zero energy. In the quantum world, even in its lowest energy state, the particle is still in motion, possessing a minimum "zero-point energy." This is a manifestation of the Heisenberg uncertainty principle.
- Probability Distributions: The square of the wave function, |ψn(x)|², gives the probability density of finding the particle at a particular position x. For the particle in a box, this probability is not uniform. For example, in the ground state (n=1), the particle is most likely to be found in the center of the box. In the first excited state (n=2), it is most likely to be found at two points, and there is a "node" in the center where the probability of finding the particle is zero.
The particle-in-a-box model, though simple, has real-world applications. It provides a basic framework for understanding the behavior of electrons in conjugated molecules and in "quantum dots," tiny semiconductor crystals whose electronic and optical properties are governed by quantum confinement.
The Quantum Harmonic Oscillator: A Universal Model
Another cornerstone of quantum mechanics is the quantum harmonic oscillator. This system describes a particle that experiences a restoring force proportional to its displacement from an equilibrium position, much like a mass on a spring. The potential energy for a classical harmonic oscillator is a parabola: V(x) = ½kx², where k is the spring constant.
The Hamiltonian for the one-dimensional quantum harmonic oscillator is:
Ĥ = - (ħ²/2m) (d²/dx²) + ½mω²x²where ω = √(k/m) is the classical angular frequency of the oscillator.
Solving the Schrödinger equation for this Hamiltonian is more complex than for the particle in a box, but it yields an equally profound set of results. One elegant method, developed by Paul Dirac, involves the use of "ladder operators" (creation and annihilation operators) that allow one to "climb" and "descend" the ladder of energy states.
The energy eigenvalues for the quantum harmonic oscillator are given by:
En = (n + ½)ħωwhere n = 0, 1, 2, ... is the quantum number.
Key features of the quantum harmonic oscillator include:
- Equally Spaced Energy Levels: Unlike the particle in a box, where the energy levels get farther apart as n increases, the energy levels of the harmonic oscillator are equally spaced, separated by a quantum of energy, ħω.
- Zero-Point Energy: Again, the ground state energy (n=0) is not zero, but is E₀ = ½ħω. This "zero-point energy" is a fundamental property of quantum systems and has observable consequences, for example, in the vibrational spectra of molecules.
- Wave Functions and Tunneling: The eigenstates of the harmonic oscillator are described by Hermite polynomials multiplied by a Gaussian function. A fascinating feature of these wave functions is that they extend into the "classically forbidden" regions, where the particle's total energy is less than the potential energy. This means there is a non-zero probability of finding the particle in a region where a classical particle could never go, a phenomenon known as "quantum tunneling."
The quantum harmonic oscillator is an incredibly important model because many physical systems, when slightly disturbed from their equilibrium positions, behave like harmonic oscillators. It is used to describe the vibrations of atoms in molecules and crystal lattices (phonons), and it forms the basis for understanding the quantization of fields in quantum field theory.
The Hydrogen Atom: The Birth of Modern Chemistry
Perhaps the most triumphant application of the Hamiltonian and the Schrödinger equation in the early days of quantum mechanics was the solution for the hydrogen atom. The hydrogen atom, consisting of a single proton and a single electron, is the simplest atom, and understanding its structure was a major goal of early 20th-century physics.
The Hamiltonian for the electron in a hydrogen atom includes the kinetic energy of the electron and the potential energy of the electrostatic (Coulomb) attraction between the negatively charged electron and the positively charged proton:
Ĥ = - (ħ²/2μ)∇² - e²/(4πε₀r)where μ is the reduced mass of the electron-proton system, ∇² is the Laplacian operator (which includes derivatives with respect to all three spatial coordinates), e is the elementary charge, ε₀ is the permittivity of free space, and r is the distance between the electron and the proton.
Solving the Schrödinger equation in three dimensions for this Hamiltonian is a mathematical tour de force, but the results are a spectacular success. The solution reveals a set of quantum numbers that precisely describe the allowed states of the electron:
- The principal quantum number, n (1, 2, 3, ...), which primarily determines the energy of the electron.
- The azimuthal or orbital angular momentum quantum number, l (0, 1, ..., n-1), which determines the shape of the electron's orbital.
- The magnetic quantum number, ml (-l, ..., 0, ..., +l), which determines the orientation of the orbital in space.
The energy eigenvalues for the hydrogen atom are given by:
En = -μe⁴ / (8ε₀²h²n²) = -13.6 eV / n²This formula perfectly matched the experimentally observed spectral lines of hydrogen, providing stunning confirmation of the new quantum theory.
The eigenstates, or wave functions, of the hydrogen atom are the familiar atomic orbitals (1s, 2s, 2p, etc.) that form the foundation of modern chemistry. The shapes of these orbitals, which describe the probability of finding the electron in a particular region of space, explain the nature of chemical bonds and the structure of the periodic table.
The solution of the hydrogen atom using the Hamiltonian was a monumental achievement. It demonstrated that quantum mechanics could not only explain the strange behavior of simple systems but could also provide a quantitative and predictive theory for the structure of matter itself.
The Hamiltonian in the Modern World: From Molecules to Quantum Computers
The power of the Hamiltonian extends far beyond the simple, idealized systems of introductory quantum mechanics. It is a workhorse of modern physics and chemistry, enabling scientists to tackle complex problems at the frontiers of knowledge.
The Hamiltonian in Quantum Chemistry
In the realm of quantum chemistry, the goal is to solve the Schrödinger equation for molecules to predict their structure, properties, and reactivity. The molecular Hamiltonian is a much more complex beast than that of the hydrogen atom. For a molecule with multiple electrons and nuclei, the Hamiltonian includes terms for:
- The kinetic energy of each electron.
- The kinetic energy of each nucleus.
- The electrostatic attraction between each electron and each nucleus.
- The electrostatic repulsion between each pair of electrons.
- The electrostatic repulsion between each pair of nuclei.
Solving the Schrödinger equation with this full Hamiltonian is an impossibly difficult task for all but the simplest molecules. To make progress, chemists employ a series of approximations, the most fundamental of which is the Born-Oppenheimer approximation. This approximation assumes that because nuclei are much heavier and move much more slowly than electrons, we can consider the nuclei to be fixed in place when solving for the motion of the electrons.
Even with this simplification, the problem of electron-electron repulsion remains a major challenge. This is where methods like Hartree-Fock theory and Density Functional Theory (DFT) come into play.
- Hartree-Fock Theory: This method approximates the complex, many-electron wave function as a single Slater determinant, which is a specific type of wave function that properly accounts for the Pauli exclusion principle (the fact that no two electrons can occupy the same quantum state). In the Hartree-Fock approximation, each electron is considered to move in an average electric field created by all the other electrons. This leads to a simplified, effective Hamiltonian for each electron, which can be solved iteratively until a self-consistent solution is found.
- Density Functional Theory (DFT): DFT takes a different approach. Instead of trying to calculate the complex many-electron wave function, DFT is based on a theorem by Hohenberg and Kohn which states that the ground state energy of a system is a unique functional of the electron density. This means that if you know the electron density (a function of only three spatial coordinates, regardless of the number of electrons), you can, in principle, calculate the exact ground state energy. In practice, the exact form of this functional is unknown, and DFT relies on approximations for the "exchange-correlation" part of the energy. The Kohn-Sham formulation of DFT introduces a fictitious system of non-interacting electrons that have the same density as the real, interacting system, and solves for the orbitals of this system using an effective Hamiltonian.
These methods, and their many refinements, have transformed chemistry, allowing for the computational design of new molecules, the study of reaction mechanisms, and the prediction of molecular properties with remarkable accuracy. At the heart of all these methods lies the Hamiltonian, the ultimate arbiter of molecular structure and energy.
The Hamiltonian in Condensed Matter Physics
In condensed matter physics, which studies the properties of solids and liquids, the Hamiltonian is used to understand the collective behavior of vast numbers of interacting particles. The goal is to explain macroscopic phenomena like conductivity, magnetism, and superconductivity from the underlying quantum mechanics of electrons in a crystal lattice.
- The Hubbard Model: One of the most important and widely studied models in condensed matter physics is the Hubbard model. It is a simplified Hamiltonian that captures the essential physics of interacting electrons in a solid. The Hubbard Hamiltonian has two main terms: a "hopping" term that allows electrons to move between adjacent lattice sites (representing kinetic energy), and an "on-site repulsion" term that exacts an energy cost U if two electrons occupy the same lattice site (representing the Coulomb repulsion between electrons).
The Hubbard model beautifully illustrates the competition between kinetic and potential energy. When the hopping term dominates, electrons are delocalized and can move freely through the material, leading to metallic behavior. When the on-site repulsion U is large, electrons become localized on individual atoms to avoid the energy penalty of double occupancy, leading to an insulating state known as a "Mott insulator." The Hubbard model is a cornerstone for understanding strongly correlated electron systems and is thought to be relevant to high-temperature superconductivity.
- The Ising Model: To describe magnetism, physicists use models like the Ising model. In this model, the "spins" of atoms on a lattice can point either "up" or "down" (+1 or -1). The Ising Hamiltonian describes the interaction between neighboring spins. In a ferromagnetic Ising model, the energy is lower when neighboring spins are aligned, leading to a state where all spins point in the same direction, creating a macroscopic magnet. The Ising model, despite its simplicity, exhibits a phase transition at a critical temperature, above which the thermal energy is sufficient to overcome the spin-spin interactions, and the material loses its magnetism.
- BCS Theory of Superconductivity: Superconductivity, the phenomenon of zero electrical resistance at low temperatures, is a macroscopic quantum effect. The Bardeen-Cooper-Schrieffer (BCS) theory explains conventional superconductivity by proposing that electrons can form bound pairs, called "Cooper pairs," due to an effective attractive interaction mediated by vibrations of the crystal lattice (phonons). The BCS Hamiltonian describes these interacting electrons and shows that the ground state of the system is a condensate of Cooper pairs. This pairing opens up an energy gap at the Fermi level, which is responsible for the remarkable properties of superconductors.
The Hamiltonian in the Age of Quantum Computing
The Hamiltonian is not just a tool for understanding the quantum world; it is also a target for simulation and a means of computation in the burgeoning field of quantum computing.
- Hamiltonian Simulation: One of the most promising applications of quantum computers is the simulation of other quantum systems. Many problems in chemistry and materials science are intractable for classical computers because the number of variables needed to describe a quantum system grows exponentially with the number of particles. Quantum computers, which operate on the principles of quantum mechanics, are naturally suited for this task. The goal of Hamiltonian simulation is to use a controllable quantum system (a quantum computer) to model the time evolution of a less controllable or more complex quantum system. This involves devising a sequence of quantum gates that approximates the evolution operator e-iĤt/ħ, where Ĥ is the Hamiltonian of the system being simulated.
- Variational Quantum Eigensolver (VQE): VQE is a hybrid quantum-classical algorithm designed to find the ground state energy of a molecule or material, which is often the most important property for understanding its behavior. The VQE algorithm uses a quantum computer to prepare a trial wave function, or "ansatz," which depends on a set of parameters. The quantum computer then measures the expectation value of the Hamiltonian for this trial state. This expectation value is then fed to a classical optimizer, which adjusts the parameters of the ansatz to minimize the energy. This process is repeated until the minimum energy—the ground state energy—is found. VQE is a leading candidate for demonstrating a practical quantum advantage in the near term.
- Adiabatic Quantum Computing (AQC): AQC is a different paradigm for quantum computation that relies on the adiabatic theorem. The process begins with a simple Hamiltonian whose ground state is easy to prepare. This initial Hamiltonian is then slowly and continuously transformed into a final, more complex Hamiltonian whose ground state encodes the solution to a computational problem. According to the adiabatic theorem, if this transformation is done slowly enough, the system will remain in its ground state throughout the evolution. At the end of the process, the system is in the ground state of the final Hamiltonian, and by measuring the state of the qubits, the solution to the problem can be read out.
The Philosophical Implications: A Clockwork Universe with a Probabilistic Heart
The central role of the Hamiltonian in quantum mechanics raises profound philosophical questions about the nature of reality. The Schrödinger equation, which governs the evolution of the wave function under the influence of the Hamiltonian, is completely deterministic. If you know the wave function of a system at a given time, the Schrödinger equation can, in principle, predict its wave function at any future time.
This deterministic evolution of the wave function seems to be at odds with the inherent randomness of quantum measurements. When we measure a quantum system, the wave function is said to "collapse" into one of its eigenstates, and the outcome of this measurement is fundamentally probabilistic. We can only predict the probability of obtaining a particular outcome, not the outcome itself.
This "measurement problem" has been a source of debate and controversy since the birth of quantum mechanics. Different interpretations of quantum mechanics offer different perspectives on this issue.
The Copenhagen interpretation, for example, posits a fundamental distinction between the deterministic evolution of the wave function and the probabilistic nature of measurement. The act of measurement is seen as a special process that is not described by the Schrödinger equation.
Other interpretations, such as the many-worlds interpretation, suggest that the Schrödinger equation is all there is. In this view, the wave function never collapses. Instead, every possible outcome of a measurement occurs in a separate, parallel universe.
Bohmian mechanics, or de Broglie-Bohm theory, is another deterministic interpretation. It postulates the existence of hidden variables—the actual positions of the particles—which evolve deterministically under the guidance of the wave function.
Regardless of which interpretation one favors, the Hamiltonian remains at the heart of the matter. It is the engine that drives the deterministic evolution of the wave function, the "clockwork" of the quantum universe. The mystery lies in how this deterministic evolution gives rise to the probabilistic world we experience through measurement.
The Hamiltonian, in a sense, represents the knowable, the deterministic aspect of the quantum world. It is the law that governs the realm of possibilities. The act of measurement, on the other hand, is the bridge between this realm of possibilities and the concrete, singular reality of our experience.
Conclusion: The Enduring Power of a Single Equation
From its origins in the elegant reformulation of classical mechanics to its central role in the quantum revolution and its continued relevance at the frontiers of modern science, the Hamiltonian has proven to be one of the most powerful and enduring concepts in all of physics. It is a testament to the power of mathematical abstraction to capture the fundamental workings of the universe.
The Hamiltonian is more than just an equation; it is a unifying principle that connects the classical and quantum worlds, the macroscopic and the microscopic, the deterministic and the probabilistic. It is the unseen architect that shapes the structure of atoms, the properties of materials, and the very evolution of the quantum state. As we continue to explore the mysteries of the quantum realm and build new technologies that harness its power, the Hamiltonian will undoubtedly remain our most trusted guide, a single equation that unlocks the secrets of a universe far stranger and more beautiful than we could have ever imagined.
Reference:
- http://materias.df.uba.ar/sa2020c2/files/2012/07/Lecture-BCS_Fernandes.pdf
- https://sinbad989.github.io/bookreading/hubbard_model_an_overview.pdf
- https://en.wikipedia.org/wiki/Hubbard_model
- https://en.wikipedia.org/wiki/Molecular_Hamiltonian
- https://pennylane.ai/topics/hamiltonian-simulation
- https://en.wikipedia.org/wiki/Adiabatic_quantum_computation
- https://qt.ntu.edu.tw/qoa/wp-content/uploads/2022/02/activity20210830stu5.pdf
- https://www.cs.umd.edu/~amchilds/talks/ibm13.pdf
- https://pubs.aip.org/aip/jcp/article/157/24/244106/2842847/Perspectives-on-determinism-in-quantum-mechanics
- https://pennylane.ai/qml/demos/tutorial_vqe/
- https://physics.stackexchange.com/questions/400162/what-exactly-is-deterministic-in-schr%C3%B6dingers-equation
- https://sufficientlywise.org/2018/03/15/determinism-and-quantum-mechanics/
- https://academic.oup.com/ptp/article/34/5/734/1857045
- https://phas.ubc.ca/~berciu/TEACHING/PHYS502/NOTES/se.pdf
- https://medium.com/quantum-untangled/hamiltonian-simulation-with-quantum-computation-fedc7cdc02e0
- https://docs.classiq.io/latest/user-guide/read/hamiltonian_simulation_guide/
- https://www.quora.com/Why-is-the-Schr%C3%B6dinger-equation-deterministic-while-measurements-are-not
- https://theory.leeds.ac.uk/interaction-distance/applications/ising/
- https://www.imperial.ac.uk/media/imperial-college/research-centres-and-groups/computational-materials-science/teaching/DFT_NATO.pdf
- https://math.columbian.gwu.edu/adiabatic-quantum-computing-construction-hamiltonian-operators
- https://www.quera.com/glossary/variational-quantum-eigensolver
- https://arxiv.org/abs/2111.05176
- https://www.cond-mat.de/events/correl15/manuscripts/mielke.pdf
- https://pubmed.ncbi.nlm.nih.gov/36586987/
- https://www.southampton.ac.uk/assets/centresresearch/documents/compchem/DFT_L2.pdf
- http://lampz.tugraz.at/~hadley/ss1/molecules/hamiltonian.php
- https://phas.ubc.ca/~berciu/TEACHING/PHYS502/NOTES/wi.pdf
- https://people.phys.ethz.ch/~ivanov/cmt/1314/ssth-lectureBC.pdf
- https://psiberg.com/hamiltonian-operator/
- https://arxiv.org/html/2303.09911v4
- https://en.wikipedia.org/wiki/Hartree%E2%80%93Fock_method
- https://web.ornl.gov/~kentpr/thesis/pkthnode13.html
- https://compphysics.github.io/ComputationalPhysics2/doc/LectureNotes/_build/html/hartreefocktheory.html
- https://web.stanford.edu/~kimth/www-mit/thk_hartreefock.pdf
- https://www.lct.jussieu.fr/pagesperso/savin/papers/paolajulien-montreal/GorTouSav-07.pdf
- https://octopus-code.org/documentation/11/manual/basics/hamiltonian/
- https://www.cond-mat.de/events/correl16/manuscripts/scalettar.pdf
- https://en.wikipedia.org/wiki/Ising_model
- https://www.thphys.uni-heidelberg.de/~wolschin/statsem21_3s.pdf
- https://www.weizmann.ac.il/condmat/oreg/sites/condmat.oreg/files/uploads/2016/tutorial_3.pdf
- https://physics.stackexchange.com/questions/186992/questions-concerning-bcs-theory-particularly-the-pairing-hamiltonian
- https://en.wikipedia.org/wiki/Variational_quantum_eigensolver
- https://www.cl.cam.ac.uk/teaching/1920/QuantComp/Quantum_Computing_Lecture_15.pdf
- https://www.scottaaronson.com/showcase2/charles-epstein.pdf