1. Introduction to many-body physics

1.1. Introduction to the course

These lecture notes aim at giving an introduction to the quantum mechanics of many-body systems and the methods relevant for many-body problems in such diverse areas as atomic, molecular, solid-state and nuclear physics, chemistry and materials science. A theoretical understanding of the behavior of quantum-mechanical many-body systems, that is, systems containing many interacting particles - is a considerable challenge in that, normally, no exact solution can be found. Instead, reliable methods are needed for approximate but accurate simulations of such systems.

These notes cover central many-body methods used in a variety of fields in physics. Starting with basic definitions and second quantization and Wick’s theorems for systems of fermions and bosons, we move on to the introduction of Feynman diagrams and many-body methods like full configuration interaction theory and various mean field approaches like Hartree-Fock theory and density functional theory. Thereafter, so-called post Hartree-Fock methods like many-body perturbation theory, coupled cluster theory (standard and unitary) and other methods will be discussed. Finally, algorithms from quantum computing for solving quantum mechanical many-body problems will be discussed. The present set of notes contain more material than covered by a regular course and some of these notes can be used for self-studies or advanced many-body topics. Depending on the interest of the participants, selected methods can be emphasized over other ones.

This chapter serves the aim of reminding the reader of some basic properties of quantum mechanics, in addition to providing a brief review of linear algebra, which lays the foundations for most of our mathematical formalism. It plays a central role in our studies and together with computational methods and various linear algebra libraries, it provides us with the tools to study complicated systems of many interacting particles. Our short reminder on linear algebra allows us to catch more than one bird with a stone, we take the liberty of introducing some basic notations and definitions used throughout these lectures.

1.2. Important Matrix and vector handling packages

Linear algebra plays a central role in many-body physics, as it does in many other mathematical descriptions of problems in physics and in Science in general. There are several central software packages for linear algebra and eigenvalue problems. Several of the more popular ones have been wrapped into ofter software packages like those from the widely used text Numerical Recipes. The original source codes in many of the available packages are often taken from the widely used software package LAPACK, which follows two other popular packages developed in the 1970s, namely EISPACK and LINPACK. We describe them shortly here.

  • LINPACK: package for linear equations and least square problems.

  • LAPACK:package for solving symmetric, unsymmetric and generalized eigenvalue problems. From LAPACK’s website http://www.netlib.org it is possible to download for free all source codes from this library. Both C/C++ and Fortran versions are available.

  • BLAS (I, II and III): (Basic Linear Algebra Subprograms) are routines that provide standard building blocks for performing basic vector and matrix operations. Blas I is vector operations, II vector-matrix operations and III matrix-matrix operations. Highly parallelized and efficient codes, all available for download from http://www.netlib.org.

When dealing with matrices and vectors a central issue is memory handling and allocation. If our code is written in Python the way we declare these objects and the way they are handled, interpreted and used by say a linear algebra library, requires codes that interface our Python program with such libraries. For Python programmers, Numpy is by now the standard Python package for numerical arrays in Python as well as the source of functions which act on these arrays. These functions span from eigenvalue solvers to functions that compute the mean value, variance or the covariance matrix. If you are not familiar with how arrays are handled in say Python or compiled languages like C++ and Fortran, the sections in this chapter may be useful. For C++ programmer, Armadillo is widely used library for linear algebra and eigenvalue problems. In addition it offers a convenient way to handle and organize arrays. We discuss this library as well. Before we proceed we believe it may be convenient to repeat some basic features of vectors, matrices, tensors etc. For systems of fermions, our ansatz for encoding the needed antisymmetry of the wave function, is to assume that we can build a many-body state based on physically relevant single-particle states. This leads the so-called Salter determinant.

1.3. Basic Matrix Features

Matrix properties reminder

\[\begin{split} \mathbf{A} = \begin{bmatrix} a_{11} & a_{12} & a_{13} & a_{14} \\ a_{21} & a_{22} & a_{23} & a_{24} \\ a_{31} & a_{32} & a_{33} & a_{34} \\ a_{41} & a_{42} & a_{43} & a_{44} \end{bmatrix}\qquad \mathbf{I} = \begin{bmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \end{bmatrix} \end{split}\]

The inverse of a matrix is defined by

\[ \mathbf{A}^{-1} \cdot \mathbf{A} = I \]
Relations Name matrix elements
$A = A^{T}$ symmetric $a_{ij} = a_{ji}$
$A = \left (A^{T} \right )^{-1}$ real orthogonal $\sum_k a_{ik} a_{jk} = \sum_k a_{ki} a_{kj} = \delta_{ij}$
$A = A^{ * }$ real matrix $a_{ij} = a_{ij}^{ * }$
$A = A^{\dagger}$ hermitian $a_{ij} = a_{ji}^{ * }$
$A = \left (A^{\dagger} \right )^{-1}$ unitary $\sum_k a_{ik} a_{jk}^{ * } = \sum_k a_{ki}^{ * } a_{kj} = \delta_{ij}$

Some famous Matrices we often encounter in many-body physics are

  • Diagonal if \(a_{ij}=0\) for \(i\ne j\)

  • Upper triangular if \(a_{ij}=0\) for \(i > j\)

  • Lower triangular if \(a_{ij}=0\) for \(i < j\)

  • Upper Hessenberg if \(a_{ij}=0\) for \(i > j+1\)

  • Lower Hessenberg if \(a_{ij}=0\) for \(i < j+1\)

  • Tridiagonal if \(a_{ij}=0\) for \(|i -j| > 1\)

  • Lower banded with bandwidth \(p\): \(a_{ij}=0\) for \(i > j+p\)

  • Upper banded with bandwidth \(p\): \(a_{ij}=0\) for \(i < j+p\)

  • Banded, block upper triangular, block lower triangular….

Furthermore, for an \(N\times N\) matrix \(\mathbf{A}\) the following properties are all equivalent

  • If the inverse of \(\mathbf{A}\) exists, \(\mathbf{A}\) is nonsingular.

  • The equation \(\mathbf{Ax}=0\) implies \(\mathbf{x}=0\).

  • The rows of \(\mathbf{A}\) form a basis of \(R^N\).

  • The columns of \(\mathbf{A}\) form a basis of \(R^N\).

  • \(\mathbf{A}\) is a product of elementary matrices.

  • \(0\) is not eigenvalue of \(\mathbf{A}\).

1.4. Numpy and arrays

Numpy provides an easy way to handle arrays in Python. The standard way to import this library is as

import numpy as np
n = 10
x = np.random.normal(size=n)
print(x)
[-0.18181242 -0.82390948 -0.95197186  0.17706887  0.80026555 -0.45625011
 -0.18799733 -0.34267102 -1.12100632 -0.18662716]

Here we have defined a vector \(x\) with \(n=10\) elements with its values given by the Normal distribution \(N(0,1)\). Another alternative is to declare a vector as follows

import numpy as np
x = np.array([1, 2, 3])
print(x)
[1 2 3]

Here we have defined a vector with three elements, with \(x_0=1\), \(x_1=2\) and \(x_2=3\). Note that both Python and C++ start numbering array elements from \(0\) and on. This means that a vector with \(n\) elements has a sequence of entities \(x_0, x_1, x_2, \dots, x_{n-1}\). We could also let (recommended) Numpy to compute the logarithms of a specific array as

import numpy as np
x = np.log(np.array([4, 7, 8]))
print(x)
[1.38629436 1.94591015 2.07944154]

Here we have used Numpy’s unary function \(np.log\). This function is highly tuned to compute array elements since the code is vectorized and does not require looping. We normaly recommend that you use the Numpy intrinsic functions instead of the corresponding log function from Python’s math module. The looping is done explicitely by the np.log function. The alternative, and slower way to compute the logarithms of a vector would be to write

import numpy as np
from math import log
x = np.array([4, 7, 8])
for i in range(0, len(x)):
    x[i] = log(x[i])
print(x)
[1 1 2]

We note that our code is much longer already and we need to import the log function from the math module. The attentive reader will also notice that the output is \([1, 1, 2]\). Python interprets automacally our numbers as integers (like the automatic keyword in C++). To change this we could define our array elements to be double precision numbers as

import numpy as np
x = np.log(np.array([4, 7, 8], dtype = np.float64))
print(x)
[1.38629436 1.94591015 2.07944154]

or simply write them as double precision numbers (Python uses 64 bits as default for floating point type variables), that is

import numpy as np
x = np.log(np.array([4.0, 7.0, 8.0]))
print(x)
[1.38629436 1.94591015 2.07944154]

To check the number of bytes (remember that one byte contains eight bits for double precision variables), you can use simple use the itemsize functionality (the array \(x\) is actually an object which inherits the functionalities defined in Numpy) as

import numpy as np
x = np.log(np.array([4.0, 7.0, 8.0]))
print(x.itemsize)
8

Having defined vectors, we are now ready to try out matrices. We can define a \(3 \times 3 \) real matrix \(\hat{A}\) as (recall that we user lowercase letters for vectors and uppercase letters for matrices)

import numpy as np
A = np.log(np.array([ [4.0, 7.0, 8.0], [3.0, 10.0, 11.0], [4.0, 5.0, 7.0] ]))
print(A)
[[1.38629436 1.94591015 2.07944154]
 [1.09861229 2.30258509 2.39789527]
 [1.38629436 1.60943791 1.94591015]]

If we use the shape function we would get \((3, 3)\) as output, that is verifying that our matrix is a \(3\times 3\) matrix. We can slice the matrix and print for example the first column (Python organized matrix elements in a row-major order, see below) as

import numpy as np
A = np.log(np.array([ [4.0, 7.0, 8.0], [3.0, 10.0, 11.0], [4.0, 5.0, 7.0] ]))
# print the first column, row-major order and elements start with 0
print(A[:,0])
[1.38629436 1.09861229 1.38629436]

We can continue this was by printing out other columns or rows. The example here prints out the second column

import numpy as np
A = np.log(np.array([ [4.0, 7.0, 8.0], [3.0, 10.0, 11.0], [4.0, 5.0, 7.0] ]))
# print the first column, row-major order and elements start with 0
print(A[1,:])
[1.09861229 2.30258509 2.39789527]

Numpy contains many other functionalities that allow us to slice, subdivide etc etc arrays. We strongly recommend that you look up the Numpy website for more details. Useful functions when defining a matrix are the np.zeros function which declares a matrix of a given dimension and sets all elements to zero

import numpy as np
n = 10
# define a matrix of dimension 10 x 10 and set all elements to zero
A = np.zeros( (n, n) )
print(A)
[[0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]
 [0. 0. 0. 0. 0. 0. 0. 0. 0. 0.]]

or initializing all elements to

import numpy as np
n = 10
# define a matrix of dimension 10 x 10 and set all elements to one
A = np.ones( (n, n) )
print(A)
[[1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]
 [1. 1. 1. 1. 1. 1. 1. 1. 1. 1.]]

or as unitarily distributed random numbers (see the material on random number generators in the statistics part)

import numpy as np
n = 10
# define a matrix of dimension 10 x 10 and set all elements to random numbers with x \in [0, 1]
A = np.random.rand(n, n)
print(A)
[[0.42668154 0.44529398 0.81431386 0.41532905 0.08538057 0.68954186
  0.05741467 0.62343179 0.29156132 0.50311603]
 [0.75838834 0.6090632  0.32329292 0.12567032 0.44684508 0.90291719
  0.67409939 0.30861483 0.27160845 0.37386191]
 [0.77853093 0.83642995 0.77942173 0.87281744 0.43413506 0.33728723
  0.82678561 0.13855613 0.7705333  0.96473269]
 [0.0632419  0.50224636 0.26817617 0.0563718  0.55243532 0.58035027
  0.46833658 0.69820786 0.98841162 0.17397314]
 [0.35748561 0.48144532 0.83970178 0.92497464 0.65877538 0.02446637
  0.40784019 0.83621124 0.47947519 0.1382501 ]
 [0.56973725 0.38465877 0.85538751 0.13002132 0.30504484 0.82358887
  0.10074632 0.02949402 0.77084258 0.93282264]
 [0.44722558 0.24534328 0.12840244 0.9423928  0.1366858  0.99570313
  0.87400321 0.69250731 0.24558877 0.04518465]
 [0.94924418 0.37864605 0.10082995 0.06187188 0.28347895 0.97877997
  0.69113578 0.01888094 0.46364549 0.42022255]
 [0.49468485 0.47782191 0.51235668 0.57404905 0.45231321 0.47933511
  0.27193391 0.15257909 0.28908355 0.19820541]
 [0.28773096 0.11602248 0.34159001 0.39539761 0.78975011 0.62034427
  0.5657688  0.32520647 0.9107702  0.80607227]]

1.5. Other Matrix and Vector Operations

The following examples show how to compute various quantities like the mean value of a matrix or a vector and how to use functions like reshape and ravel. These are all useful quantities when scaling the data and preparing the data for various machine learning algorithms and when calculating quantities like the mean squared error or the variance.

"""
Simple code that tests various numpy functions
"""

import numpy as np
# Simple test-matrix of dim 3 x 4
a = np.array([ [1, 2, 3], [4, 5, 6], [7, 8, 9],[10, 11, 12]],dtype=np.float64)
print(f"The test matrix:{a}")
# This is the total mean summed over all elements, which here has to be 6.5
print(f"This is the total mean summed over all elements:{np.mean(a,dtype=np.float64)}")
# This is the mean for each column, it returns an array with the mean values for each column. It returns a row-like vector
print(f"This is the mean for each column:{np.mean(a, axis=0, keepdims=True,dtype=np.float64)}")
# This is the mean value for each row, it returns an array via the keepdims option which is a column-like vector if
# keepdims=True. Else it return a row-like vector
# Try setting keepdims=False
print(f"This is the mean value for each row:{np.mean(a, axis=1, keepdims=True,dtype=np.float64)}")
# We print then the mean value for each row by  setting keepdims=False
print(f"This is the mean value for each row with keepdims false:{np.mean(a, axis=1, keepdims=False,dtype=np.float64)}")

# Ravel return a contiguous flattened array.
print(f"Flatten  the matrix:{np.ravel(a)}")
# It is the same as reshaping the matrix into a one-dimensional array
print(f"Reshape the matrix to a one-dim array:{a.reshape(-1)}")
#  ‘C’ means to index the elements in row-major, C-style order, with the last axis index changing fastest, back to the first axis index changing slowest.
# ‘F’ means to index the elements in column-major, Fortran-style order, with the first index changing fastest, and the last index changing slowest 
print(np.ravel(a, order='F'))
# When order is ‘A’, it will preserve the array’s ‘C’ or ‘F’ ordering
# ‘A’ means to read the elements in Fortran-like index order if a is Fortran contiguous in memory, C-like order otherwise.
# ‘K’ means to read the elements in the order they occur in memory, except for reversing the data when strides are negative. By default, ‘C’ index order is used.
# Transposing it
print(np.ravel(a.T))
print(np.ravel(a.T, order='A'))
The test matrix:[[ 1.  2.  3.]
 [ 4.  5.  6.]
 [ 7.  8.  9.]
 [10. 11. 12.]]
This is the total mean summed over all elements:6.5
This is the mean for each column:[[5.5 6.5 7.5]]
This is the mean value for each row:[[ 2.]
 [ 5.]
 [ 8.]
 [11.]]
This is the mean value for each row with keepdims false:[ 2.  5.  8. 11.]
Flatten  the matrix:[ 1.  2.  3.  4.  5.  6.  7.  8.  9. 10. 11. 12.]
Reshape the matrix to a one-dim array:[ 1.  2.  3.  4.  5.  6.  7.  8.  9. 10. 11. 12.]
[ 1.  4.  7. 10.  2.  5.  8. 11.  3.  6.  9. 12.]
[ 1.  4.  7. 10.  2.  5.  8. 11.  3.  6.  9. 12.]
[ 1.  2.  3.  4.  5.  6.  7.  8.  9. 10. 11. 12.]

1.6. Definitions and vector and matrix operations

We use this section to define some central quantities as well as presenting our notations for single-particle and many-body states. We will normally tend to label vectors and states with the so-called Dirac bra-ket notation. In general we will use lower-case and boldfaced letters for vectors and upper-case boldfaced letters for matrices. These ways to represent vectors will however be replaced with the above-mentioned Dirac bra-ket notation. As example, consider the vector \(\boldsymbol{x}\) defined here

\[\begin{split} \boldsymbol{x} = \vert x \rangle = \begin{bmatrix} x_0 \\ x_1 \\ x_2 \\ \dots \\ x_{n-2} \\ x_{n-1} \end{bmatrix} \in \mathbb{C}^{n} , \end{split}\]

and its hermitian conjugate

\[ \langle x \vert = \vert x \rangle^{\dagger}=\begin{bmatrix} x_0^* & x_1^* & x_2^* & \dots & x_{n-2}^* & x_{n-1}^* \end{bmatrix}. \]

The norm of two vectors is defined as (assuming they are not orthogonal

\[ \langle x \vert y \rangle =\sum_ix_i^*y_i, \]

and for an orthonormal (normalized and orthogonal) basis we have

\[ \langle x_j \vert x_i \rangle =\delta_{ij}. \]

We can also use this orthonormal basis to define the completeness relation as

\[ \boldsymbol{I}=\sum_{i=0}^{\infty}\vert x_i \rangle \langle _i \vert, \]

where we have defined the outer product

\[\begin{split} \vert x \rangle \langle y \vert = \begin{bmatrix} x_0y_0^* & x_0y_1^* & x_0y_2^* & \dots & x_0y_{n-2}^* & x_0y_{n-1}^* \\ x_1y_0^* & x_1y_1^* & x_1y_2^* & \dots & x_1y_{n-2}^* & x_1y_{n-1}^* \\ \dots & \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots & \dots \\ x_{n-2}y_0^* & x_{n-2}y_1^* & x_{n-2}y_2^* & \dots & x_{n-2}y_{n-2}^* & x_{n-2}y_{n-1}^* \\ x_{n-1}y_0^* & x_{n-1}y_1^* & x_{n-1}y_2^* & \dots & x_{n-1}y_{n-2}^* & x_{n-1}y_{n-1}^* \\ \end{bmatrix}. \end{split}\]

The outer product defined here is going to play an important role in these studies as it can be used to define important projection operators. Since the basis of orthonormal states is, normally, infinite, the above sum runs over an infinite basis. However, for all practical reasons, unless we can find analytical solutions to a quantum mechanical problem, the above sum needs to be truncated.

An orthonormal basis can in turn be used to define a new basis set \(\vert \phi_i \rangle\)

\[ \vert \phi_i \rangle \sum_j C_{ij} \vert x_j\rangle, \]

where the coefficients \(C_{ij}=\langle \phi_i\vert x_j\rangle\) are the matrix elements of a transformation from a basis set \(\vert \boldsymbol{x} \rangle\) to a new basis set \(\vert \boldsymbol{\phi}\rangle\). The matrix \(\boldsymbol{C}\) is normally a unitary matrix.

With the outer product we can define projection operators \(\hat{P}\) and \(\hat{Q}\). Consider for example the two orthonormal states

\[\begin{split} \vert x_0 \rangle = \begin{bmatrix} 1 \\ 0 \end{bmatrix}, \end{split}\]

and

\[\begin{split} \vert x_1 \rangle = \begin{bmatrix} 0 \\ 1 \end{bmatrix}. \end{split}\]

We can easily define a new basis which is a linear combination of the two

\[ \vert \phi \rangle = \alpha \vert x_0 \rangle +\beta \vert x_1 \rangle \]

1.6.1. Tensor products

Consider now two vectors with length \(n=2\), with elements

\[\begin{split} \vert x \rangle = \begin{bmatrix} x_0 \\ x_1 \end{bmatrix}, \end{split}\]

and

\[\begin{split} \vert x \rangle = \begin{bmatrix} y_0 \\ y_1 \end{bmatrix}, \end{split}\]
\[\begin{split} \vert x \rangle \otimes \vert y \rangle = \vert xy \rangle = \begin{bmatrix} x_0y_0 \\ x_0y_1 \\ x_1y_0 \\ x_1y_1 \end{bmatrix}, \end{split}\]

which is now a vector of length \(4\). The tensor product of two matrices is defined as

1.6.2. Representation of states

Before we proceed we need several definitions. Throughout these lectures we will assume that the interacting part of the Hamiltonian can be approximated by a two-body interaction. This means that our Hamiltonian can be written as the sum of a onebody part, which includes kinetic energy and an eventual external field, and a twobody interaction

\[ \begin{equation} \hat{H} = \hat{H}_0 + \hat{H}_I = \sum_{i=1}^A \hat{h}_0(x_i) + \sum_{i < j}^A \hat{v}(r_{ij}), \label{Hnuclei} \tag{1} \end{equation} \]

with

\[ \begin{equation} H_0=\sum_{i=1}^A \hat{h}_0(x_i). \label{hinuclei} \tag{2} \end{equation} \]

The onebody part \(u_{\mathrm{ext}}(x_i)\) is normally approximated by a harmonic oscillator potential or the Coulomb interaction an electron feels from the nucleus. However, other potentials are fully possible, such as one derived from the self-consistent solution of the Hartree-Fock equations to be discussed here.

Our Hamiltonian is invariant under the permutation (interchange) of two particles. Since we deal with fermions however, the total wave function is antisymmetric. Let \(\hat{P}\) be an operator which interchanges two particles. Due to the symmetries we have ascribed to our Hamiltonian, this operator commutes with the total Hamiltonian,

\[ [\hat{H},\hat{P}] = 0, \]

meaning that \(\Psi_{\lambda}(x_1, x_2, \dots , x_A)\) is an eigenfunction of \(\hat{P}\) as well, that is

\[ \hat{P}_{ij}\Psi_{\lambda}(x_1, x_2, \dots,x_i,\dots,x_j,\dots,x_A)= \beta\Psi_{\lambda}(x_1, x_2, \dots,x_i,\dots,x_j,\dots,x_A), \]

where \(\beta\) is the eigenvalue of \(\hat{P}\). We have introduced the suffix \(ij\) in order to indicate that we permute particles \(i\) and \(j\). The Pauli principle tells us that the total wave function for a system of fermions has to be antisymmetric, resulting in the eigenvalue \(\beta = -1\).

In our case we assume that we can approximate the exact eigenfunction with a Slater determinant

\[\begin{split} \begin{equation} \Phi(x_1, x_2,\dots ,x_A,\alpha,\beta,\dots, \sigma)=\frac{1}{\sqrt{A!}} \left| \begin{array}{ccccc} \psi_{\alpha}(x_1)& \psi_{\alpha}(x_2)& \dots & \dots & \psi_{\alpha}(x_A)\\ \psi_{\beta}(x_1)&\psi_{\beta}(x_2)& \dots & \dots & \psi_{\beta}(x_A)\\ \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots \\ \psi_{\sigma}(x_1)&\psi_{\sigma}(x_2)& \dots & \dots & \psi_{\sigma}(x_A)\end{array} \right|, \label{eq:HartreeFockDet} \tag{3} \end{equation} \end{split}\]

where \(x_i\) stand for the coordinates and spin values of a particle \(i\) and \(\alpha,\beta,\dots, \gamma\) are quantum numbers needed to describe remaining quantum numbers.

Since we will deal with Fermions (identical and indistinguishable particles) we will form an ansatz for a given state in terms of so-called Slater determinants determined by a chosen basis of single-particle functions.

For a given \(n\times n\) matrix \(\mathbf{A}\) we can write its determinant

\[\begin{split} det(\mathbf{A})=|\mathbf{A}|= \left| \begin{array}{ccccc} a_{11}& a_{12}& \dots & \dots & a_{1n}\\ a_{21}&a_{22}& \dots & \dots & a_{2n}\\ \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots \\ a_{n1}& a_{n2}& \dots & \dots & a_{nn}\end{array} \right|, \end{split}\]

in a more compact form as

\[ |\mathbf{A}|= \sum_{i=1}^{n!}(-1)^{p_i}\hat{P}_i a_{11}a_{22}\dots a_{nn}, \]

where \(\hat{P}_i\) is a permutation operator which permutes the column indices \(1,2,3,\dots,n\) and the sum runs over all \(n!\) permutations. The quantity \(p_i\) represents the number of transpositions of column indices that are needed in order to bring a given permutation back to its initial ordering, in our case given by \(a_{11}a_{22}\dots a_{nn}\) here.

A simple \(2\times 2\) determinant illustrates this. We have

\[\begin{split} det(\mathbf{A})= \left| \begin{array}{cc} a_{11}& a_{12}\\ a_{21}&a_{22}\end{array} \right|= (-1)^0a_{11}a_{22}+(-1)^1a_{12}a_{21}, \end{split}\]

where in the last term we have interchanged the column indices \(1\) and \(2\). The natural ordering we have chosen is \(a_{11}a_{22}\).

The single-particle function \(\psi_{\alpha}(x_i)\) are eigenfunctions of the onebody Hamiltonian \(h_i\), that is

\[ \hat{h}_0(x_i)=\hat{t}(x_i) + \hat{u}_{\mathrm{ext}}(x_i), \]

with eigenvalues

\[ \hat{h}_0(x_i) \psi_{\alpha}(x_i)=\left(\hat{t}(x_i) + \hat{u}_{\mathrm{ext}}(x_i)\right)\psi_{\alpha}(x_i)=\varepsilon_{\alpha}\psi_{\alpha}(x_i). \]

The energies \(\varepsilon_{\alpha}\) are the so-called non-interacting single-particle energies, or unperturbed energies. The total energy is in this case the sum over all single-particle energies, if no two-body or more complicated many-body interactions are present.

Let us denote the ground state energy by \(E_0\). According to the variational principle we have

\[ E_0 \le E[\Phi] = \int \Phi^*\hat{H}\Phi d\mathbf{\tau} \]

where \(\Phi\) is a trial function which we assume to be normalized

\[ \int \Phi^*\Phi d\mathbf{\tau} = 1, \]

where we have used the shorthand \(d\mathbf{\tau}=dx_1dr_2\dots dr_A\).

In the Hartree-Fock method the trial function is the Slater determinant of Eq. (3) which can be rewritten as

\[ \Phi(x_1,x_2,\dots,x_A,\alpha,\beta,\dots,\nu) = \frac{1}{\sqrt{A!}}\sum_{P} (-)^P\hat{P}\psi_{\alpha}(x_1) \psi_{\beta}(x_2)\dots\psi_{\nu}(x_A)=\sqrt{A!}\hat{A}\Phi_H, \]

where we have introduced the antisymmetrization operator \(\hat{A}\) defined by the summation over all possible permutations of two particles.

It is defined as

\[ \begin{equation} \hat{A} = \frac{1}{A!}\sum_{p} (-)^p\hat{P}, \label{antiSymmetryOperator} \tag{4} \end{equation} \]

with \(p\) standing for the number of permutations. We have introduced for later use the so-called Hartree-function, defined by the simple product of all possible single-particle functions

\[ \Phi_H(x_1,x_2,\dots,x_A,\alpha,\beta,\dots,\nu) = \psi_{\alpha}(x_1) \psi_{\beta}(x_2)\dots\psi_{\nu}(x_A). \]

Both \(\hat{H}_0\) and \(\hat{H}_I\) are invariant under all possible permutations of any two particles and hence commute with \(\hat{A}\)

\[ \begin{equation} [H_0,\hat{A}] = [H_I,\hat{A}] = 0. \label{commutionAntiSym} \tag{5} \end{equation} \]

Furthermore, \(\hat{A}\) satisfies

\[ \begin{equation} \hat{A}^2 = \hat{A}, \label{AntiSymSquared} \tag{6} \end{equation} \]

since every permutation of the Slater determinant reproduces it.

The expectation value of \(\hat{H}_0\)

\[ \int \Phi^*\hat{H}_0\Phi d\mathbf{\tau} = A! \int \Phi_H^*\hat{A}\hat{H}_0\hat{A}\Phi_H d\mathbf{\tau} \]

is readily reduced to

\[ \int \Phi^*\hat{H}_0\Phi d\mathbf{\tau} = A! \int \Phi_H^*\hat{H}_0\hat{A}\Phi_H d\mathbf{\tau}, \]

where we have used Eqs. (5) and (6). The next step is to replace the antisymmetrization operator by its definition and to replace \(\hat{H}_0\) with the sum of one-body operators

\[ \int \Phi^*\hat{H}_0\Phi d\mathbf{\tau} = \sum_{i=1}^A \sum_{p} (-)^p\int \Phi_H^*\hat{h}_0\hat{P}\Phi_H d\mathbf{\tau}. \]

The integral vanishes if two or more particles are permuted in only one of the Hartree-functions \(\Phi_H\) because the individual single-particle wave functions are orthogonal. We obtain then

\[ \int \Phi^*\hat{H}_0\Phi d\mathbf{\tau}= \sum_{i=1}^A \int \Phi_H^*\hat{h}_0\Phi_H d\mathbf{\tau}. \]

Orthogonality of the single-particle functions allows us to further simplify the integral, and we arrive at the following expression for the expectation values of the sum of one-body Hamiltonians

\[ \begin{equation} \int \Phi^*\hat{H}_0\Phi d\mathbf{\tau} = \sum_{\mu=1}^A \int \psi_{\mu}^*(x)\hat{h}_0\psi_{\mu}(x)dx d\mathbf{r}. \label{H1Expectation} \tag{7} \end{equation} \]

We introduce the following shorthand for the above integral

\[ \langle \mu | \hat{h}_0 | \mu \rangle = \int \psi_{\mu}^*(x)\hat{h}_0\psi_{\mu}(x)dx, \]

and rewrite Eq. (7) as

\[ \begin{equation} \int \Phi^*\hat{H}_0\Phi d\tau = \sum_{\mu=1}^A \langle \mu | \hat{h}_0 | \mu \rangle. \label{H1Expectation1} \tag{8} \end{equation} \]

The expectation value of the two-body part of the Hamiltonian is obtained in a similar manner. We have

\[ \int \Phi^*\hat{H}_I\Phi d\mathbf{\tau} = A! \int \Phi_H^*\hat{A}\hat{H}_I\hat{A}\Phi_H d\mathbf{\tau}, \]

which reduces to

\[ \int \Phi^*\hat{H}_I\Phi d\mathbf{\tau} = \sum_{i\le j=1}^A \sum_{p} (-)^p\int \Phi_H^*\hat{v}(r_{ij})\hat{P}\Phi_H d\mathbf{\tau}, \]

by following the same arguments as for the one-body Hamiltonian.

Because of the dependence on the inter-particle distance \(r_{ij}\), permutations of any two particles no longer vanish, and we get

\[ \int \Phi^*\hat{H}_I\Phi d\mathbf{\tau} = \sum_{i < j=1}^A \int \Phi_H^*\hat{v}(r_{ij})(1-P_{ij})\Phi_H d\mathbf{\tau}. \]

where \(P_{ij}\) is the permutation operator that interchanges particle \(i\) and particle \(j\). Again we use the assumption that the single-particle wave functions are orthogonal.

We obtain

\[ \begin{equation} \int \Phi^*\hat{H}_I\Phi d\mathbf{\tau} = \frac{1}{2}\sum_{\mu=1}^A\sum_{\nu=1}^A \left[ \int \psi_{\mu}^*(x_i)\psi_{\nu}^*(x_j)\hat{v}(r_{ij})\psi_{\mu}(x_i)\psi_{\nu}(x_j) dx_idx_j \right. \label{_auto1} \tag{9} \end{equation} \]
\[ \begin{equation} \left. - \int \psi_{\mu}^*(x_i)\psi_{\nu}^*(x_j) \hat{v}(r_{ij})\psi_{\nu}(x_i)\psi_{\mu}(x_j) dx_idx_j \right]. \label{H2Expectation} \tag{10} \end{equation} \]

The first term is the so-called direct term. It is frequently also called the Hartree term, while the second is due to the Pauli principle and is called the exchange term or just the Fock term. The factor \(1/2\) is introduced because we now run over all pairs twice.

The last equation allows us to introduce some further definitions.
The single-particle wave functions \(\psi_{\mu}(x)\), defined by the quantum numbers \(\mu\) and \(x\) are defined as the overlap

\[ \psi_{\alpha}(x) = \langle x | \alpha \rangle . \]

We introduce the following shorthands for the above two integrals

\[ \langle \mu\nu|\hat{v}|\mu\nu\rangle = \int \psi_{\mu}^*(x_i)\psi_{\nu}^*(x_j)\hat{v}(r_{ij})\psi_{\mu}(x_i)\psi_{\nu}(x_j) dx_idx_j, \]

and

\[ \langle \mu\nu|\hat{v}|\nu\mu\rangle = \int \psi_{\mu}^*(x_i)\psi_{\nu}^*(x_j) \hat{v}(r_{ij})\psi_{\nu}(x_i)\psi_{\mu}(x_j) dx_idx_j. \]

1.7. Preparing for later studies: varying the coefficients of a wave function expansion and orthogonal transformations

It is common to expand the single-particle functions in a known basis and vary the coefficients, that is, the new single-particle wave function is written as a linear expansion in terms of a fixed chosen orthogonal basis (for example the well-known harmonic oscillator functions or the hydrogen-like functions etc). We define our new single-particle basis (this is a normal approach for Hartree-Fock theory) by performing a unitary transformation on our previous basis (labelled with greek indices) as

\[ \begin{equation} \psi_p^{new} = \sum_{\lambda} C_{p\lambda}\phi_{\lambda}. \label{eq:newbasis} \tag{11} \end{equation} \]

In this case we vary the coefficients \(C_{p\lambda}\). If the basis has infinitely many solutions, we need to truncate the above sum. We assume that the basis \(\phi_{\lambda}\) is orthogonal.

It is normal to choose a single-particle basis defined as the eigenfunctions of parts of the full Hamiltonian. The typical situation consists of the solutions of the one-body part of the Hamiltonian, that is we have

\[ \hat{h}_0\phi_{\lambda}=\epsilon_{\lambda}\phi_{\lambda}. \]

The single-particle wave functions \(\phi_{\lambda}(\mathbf{r})\), defined by the quantum numbers \(\lambda\) and \(\mathbf{r}\) are defined as the overlap

\[ \phi_{\lambda}(\mathbf{r}) = \langle \mathbf{r} | \lambda \rangle . \]

In deriving the Hartree-Fock equations, we will expand the single-particle functions in a known basis and vary the coefficients, that is, the new single-particle wave function is written as a linear expansion in terms of a fixed chosen orthogonal basis (for example the well-known harmonic oscillator functions or the hydrogen-like functions etc).

We stated that a unitary transformation keeps the orthogonality. To see this consider first a basis of vectors \(\mathbf{v}_i\),

\[\begin{split} \mathbf{v}_i = \begin{bmatrix} v_{i1} \\ \dots \\ \dots \\v_{in} \end{bmatrix} \end{split}\]

We assume that the basis is orthogonal, that is

\[ \mathbf{v}_j^T\mathbf{v}_i = \delta_{ij}. \]

An orthogonal or unitary transformation

\[ \mathbf{w}_i=\mathbf{U}\mathbf{v}_i, \]

preserves the dot product and orthogonality since

\[ \mathbf{w}_j^T\mathbf{w}_i=(\mathbf{U}\mathbf{v}_j)^T\mathbf{U}\mathbf{v}_i=\mathbf{v}_j^T\mathbf{U}^T\mathbf{U}\mathbf{v}_i= \mathbf{v}_j^T\mathbf{v}_i = \delta_{ij}. \]

This means that if the coefficients \(C_{p\lambda}\) belong to a unitary or orthogonal trasformation (using the Dirac bra-ket notation)

\[ \vert p\rangle = \sum_{\lambda} C_{p\lambda}\vert\lambda\rangle, \]

orthogonality is preserved, that is \(\langle \alpha \vert \beta\rangle = \delta_{\alpha\beta}\) and \(\langle p \vert q\rangle = \delta_{pq}\).

This propertry is extremely useful when we build up a basis of many-body Stater determinant based states.

Note also that although a basis \(\vert \alpha\rangle\) contains an infinity of states, for practical calculations we have always to make some truncations.

Before we develop for example the Hartree-Fock equations, there is another very useful property of determinants that we will use both in connection with Hartree-Fock calculations and later shell-model calculations.

Consider the following determinant

\[\begin{split} \left| \begin{array}{cc} \alpha_1b_{11}+\alpha_2sb_{12}& a_{12}\\ \alpha_1b_{21}+\alpha_2b_{22}&a_{22}\end{array} \right|=\alpha_1\left|\begin{array}{cc} b_{11}& a_{12}\\ b_{21}&a_{22}\end{array} \right|+\alpha_2\left| \begin{array}{cc} b_{12}& a_{12}\\b_{22}&a_{22}\end{array} \right| \end{split}\]

We can generalize this to an \(n\times n\) matrix and have

\[\begin{split} \left| \begin{array}{cccccc} a_{11}& a_{12} & \dots & \sum_{k=1}^n c_k b_{1k} &\dots & a_{1n}\\ a_{21}& a_{22} & \dots & \sum_{k=1}^n c_k b_{2k} &\dots & a_{2n}\\ \dots & \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots & \dots \\ a_{n1}& a_{n2} & \dots & \sum_{k=1}^n c_k b_{nk} &\dots & a_{nn}\end{array} \right|= \sum_{k=1}^n c_k\left| \begin{array}{cccccc} a_{11}& a_{12} & \dots & b_{1k} &\dots & a_{1n}\\ a_{21}& a_{22} & \dots & b_{2k} &\dots & a_{2n}\\ \dots & \dots & \dots & \dots & \dots & \dots\\ \dots & \dots & \dots & \dots & \dots & \dots\\ a_{n1}& a_{n2} & \dots & b_{nk} &\dots & a_{nn}\end{array} \right| . \end{split}\]

This is a property we will use in our Hartree-Fock discussions.

We can generalize the previous results, now with all elements \(a_{ij}\) being given as functions of linear combinations of various coefficients \(c\) and elements \(b_{ij}\),

\[\begin{split} \left| \begin{array}{cccccc} \sum_{k=1}^n b_{1k}c_{k1}& \sum_{k=1}^n b_{1k}c_{k2} & \dots & \sum_{k=1}^n b_{1k}c_{kj} &\dots & \sum_{k=1}^n b_{1k}c_{kn}\\ \sum_{k=1}^n b_{2k}c_{k1}& \sum_{k=1}^n b_{2k}c_{k2} & \dots & \sum_{k=1}^n b_{2k}c_{kj} &\dots & \sum_{k=1}^n b_{2k}c_{kn}\\ \dots & \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots &\dots \\ \sum_{k=1}^n b_{nk}c_{k1}& \sum_{k=1}^n b_{nk}c_{k2} & \dots & \sum_{k=1}^n b_{nk}c_{kj} &\dots & \sum_{k=1}^n b_{nk}c_{kn}\end{array} \right|=det(\mathbf{C})det(\mathbf{B}), \end{split}\]

where \(det(\mathbf{C})\) and \(det(\mathbf{B})\) are the determinants of \(n\times n\) matrices with elements \(c_{ij}\) and \(b_{ij}\) respectively.
This is a property we will use in our Hartree-Fock discussions. Convince yourself about the correctness of the above expression by setting \(n=2\).

With our definition of the new basis in terms of an orthogonal basis we have

\[ \psi_p(x) = \sum_{\lambda} C_{p\lambda}\phi_{\lambda}(x). \]

If the coefficients \(C_{p\lambda}\) belong to an orthogonal or unitary matrix, the new basis is also orthogonal. Our Slater determinant in the new basis \(\psi_p(x)\) is written as

\[\begin{split} \frac{1}{\sqrt{A!}} \left| \begin{array}{ccccc} \psi_{p}(x_1)& \psi_{p}(x_2)& \dots & \dots & \psi_{p}(x_A)\\ \psi_{q}(x_1)&\psi_{q}(x_2)& \dots & \dots & \psi_{q}(x_A)\\ \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots \\ \psi_{t}(x_1)&\psi_{t}(x_2)& \dots & \dots & \psi_{t}(x_A)\end{array} \right|=\frac{1}{\sqrt{A!}} \left| \begin{array}{ccccc} \sum_{\lambda} C_{p\lambda}\phi_{\lambda}(x_1)& \sum_{\lambda} C_{p\lambda}\phi_{\lambda}(x_2)& \dots & \dots & \sum_{\lambda} C_{p\lambda}\phi_{\lambda}(x_A)\\ \sum_{\lambda} C_{q\lambda}\phi_{\lambda}(x_1)&\sum_{\lambda} C_{q\lambda}\phi_{\lambda}(x_2)& \dots & \dots & \sum_{\lambda} C_{q\lambda}\phi_{\lambda}(x_A)\\ \dots & \dots & \dots & \dots & \dots \\ \dots & \dots & \dots & \dots & \dots \\ \sum_{\lambda} C_{t\lambda}\phi_{\lambda}(x_1)&\sum_{\lambda} C_{t\lambda}\phi_{\lambda}(x_2)& \dots & \dots & \sum_{\lambda} C_{t\lambda}\phi_{\lambda}(x_A)\end{array} \right|, \end{split}\]

which is nothing but \(det(\mathbf{C})det(\Phi)\), with \(det(\Phi)\) being the determinant given by the basis functions \(\phi_{\lambda}(x)\).

In our discussions hereafter we will use our definitions of single-particle states above and below the Fermi (\(F\)) level given by the labels \(ijkl\dots \le F\) for so-called single-hole states and \(abcd\dots > F\) for so-called particle states. For general single-particle states we employ the labels \(pqrs\dots\).

The energy functional is

\[ E[\Phi] = \sum_{\mu=1}^A \langle \mu | h | \mu \rangle + \frac{1}{2}\sum_{{\mu}=1}^A\sum_{{\nu}=1}^A \langle \mu\nu|\hat{v}|\mu\nu\rangle_{AS}, \]

we found the expression for the energy functional in terms of the basis function \(\phi_{\lambda}(\mathbf{r})\). We then varied the above energy functional with respect to the basis functions \(|\mu \rangle\). Now we are interested in defining a new basis defined in terms of a chosen basis as defined in Eq. (11). We can then rewrite the energy functional as

\[ \begin{equation} E[\Phi^{New}] = \sum_{i=1}^A \langle i | h | i \rangle + \frac{1}{2}\sum_{ij=1}^A\langle ij|\hat{v}|ij\rangle_{AS}, \label{FunctionalEPhi2} \tag{12} \end{equation} \]

where \(\Phi^{New}\) is the new Slater determinant defined by the new basis of Eq. (11).

Using Eq. (11) we can rewrite Eq. (12) as

\[ \begin{equation} E[\Psi] = \sum_{i=1}^A \sum_{\alpha\beta} C^*_{i\alpha}C_{i\beta}\langle \alpha | h | \beta \rangle + \frac{1}{2}\sum_{ij=1}^A\sum_{{\alpha\beta\gamma\delta}} C^*_{i\alpha}C^*_{j\beta}C_{i\gamma}C_{j\delta}\langle \alpha\beta|\hat{v}|\gamma\delta\rangle_{AS}. \label{FunctionalEPhi3} \tag{13} \end{equation} \]