summaryrefslogtreecommitdiff
path: root/source/know/concept/jellium/index.md
blob: 6e395b47b30112e708dbdcdc0232530cc8cd29cb (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
---
title: "Jellium"
date: 2021-11-23
categories:
- Physics
- Quantum mechanics
- Perturbation
layout: "concept"
---

**Jellium**, also called the **uniform** or **homogeneous electron gas**,
is a theoretical material where all electrons are free,
and the ions' positive charge is smeared into a uniform background "jelly".
This simple model lets us study electron interactions easily.


## Without interactions

Let us start by neglecting electron-electron interactions.
This is clearly a dubious assumption, but we will stick with it for now.
For an infinitely large sample of jellium,
the single-electron states are simply plane waves.
We consider an arbitrary cube of volume $V$,
and impose periodic boundary conditions on it,
such that the single-particle orbitals are (suppressing spin):

$$\begin{aligned}
    \Inprod{\vb{r}}{\psi_{\vb{k}}}
    = \psi_{\vb{k}}(\vb{r})
    = \frac{1}{\sqrt{V}} \exp(i \vb{k} \cdot \vb{r})
    \qquad \quad
    \vb{k} = \frac{2 \pi}{V^{1/3}} (n_x, n_y, n_z)
\end{aligned}$$

Where $n_x, n_y, n_z \in \mathbb{Z}$.
This is a discrete (but infinite) set of independent orbitals,
so it is natural to use the
[second quantization](/know/concept/second-quantization/)
to write the non-interacting Hamiltonian $\hat{H}_0$,
where $\hbar^2 |\vb{k}|^2 / (2 m)$ is the kinetic energy
of the orbital with wavevector $\vb{k}$, and $s$ is the spin:

$$\begin{aligned}
    \hat{H}_0
    = \sum_{s} \sum_{\vb{k}} \frac{\hbar^2 |\vb{k}|^2}{2 m} \hat{c}_{s,\vb{k}}^\dagger \hat{c}_{s,\vb{k}}
\end{aligned}$$

Assuming that the temperature $T = 0$,
the $N$-electron ground state of this Hamiltonian
is known as the **Fermi sea** or **Fermi sphere** $\Ket{\mathrm{FS}}$,
and is constructed by filling up the single-electron states
starting from the lowest energy:

$$\begin{aligned}
    \Ket{\mathrm{FS}}
    = \prod_{s} \prod_{j = 1}^{N/2} \hat{c}_{s,\vb{k}_j}^\dagger \Ket{0}
\end{aligned}$$

Because $T = 0$, all the electrons stay in their assigned state.
The energy and wavenumber $|\vb{k}|$ of the highest filled orbital
are called the **Fermi energy** $\epsilon_F$ and **Fermi wavenumber** $k_F$,
and obey the expected kinetic energy relation:

$$\begin{aligned}
    \boxed{
        \epsilon_F
        = \frac{\hbar^2}{2 m} k_F^2
    }
\end{aligned}$$

The Fermi sea can be visualized in $\vb{k}$-space as a sphere with radius $k_F$.
Because $\vb{k}$ is discrete, the sphere's surface is not smooth,
but in the limit $V \to \infty$ it becomes perfect.

Now, we would like a relation between the system's parameters,
e.g. $N$ and $V$, and the resulting values of $\epsilon_F$ or $k_F$.
The total population $N$ must be given by:

$$\begin{aligned}
    N
    = \sum_{s} \sum_{\vb{k}} \matrixel{\mathrm{FS}}{\hat{c}_{s,\vb{k}}^\dagger \hat{c}_{s,\vb{k}}}{\mathrm{FS}}
    = \sum_{s} \frac{V}{(2 \pi)^3} \int_{-\infty}^\infty \matrixel{\mathrm{FS}}{\hat{c}_{s,\vb{k}}^\dagger \hat{c}_{s,\vb{k}}}{\mathrm{FS}} \dd{\vb{k}}
\end{aligned}$$

Where we have turned the sum over $\vb{k}$ into an integral with a constant factor,
by using that each orbital exclusively occupies a volume $(2 \pi)^3 / V$ in $\vb{k}$-space.

At zero temperature, this inner product can only be $0$ or $1$,
depending on whether $\vb{k}$ is outside or inside the Fermi sphere.
We can therefore rewrite using a
[Heaviside step function](/know/concept/heaviside-step-function/):

$$\begin{aligned}
    N
    = \sum_{s} \frac{V}{(2 \pi)^3} \int_{-\infty}^\infty \Theta(k_F - |\vb{k}|) \dd{\vb{k}}
    = 2 \frac{V}{(2 \pi)^3} \int_{-\infty}^\infty \Theta(k_F - |\vb{k}|) \dd{\vb{k}}
\end{aligned}$$

Where we realized that spin does not matter,
and replaced the sum over $s$ by a factor $2$.
In order to evaluate this 3D integral,
we go to [spherical coordinates](/know/concept/spherical-coordinates/)
$(|\vb{k}|, \theta, \varphi)$:

$$\begin{aligned}
    N
    &= \frac{V}{4 \pi^3} \int_0^{2 \pi} \int_0^\pi \int_0^\infty \Theta(k_F - |\vb{k}|) |\vb{k}|^2 \sin(\theta) \dd{|\vb{k}|} \dd{\theta} \dd{\varphi}
    \\
    &= \frac{V}{4 \pi^3} 4 \pi \int_0^{k_F} |\vb{k}|^2 \dd{|\vb{k}|}
    = \frac{V}{\pi^2} \bigg[ \frac{|\vb{k}|^3}{3} \bigg]_0^{k_F}
    = \frac{V}{3 \pi^2} k_F^3
\end{aligned}$$

Using that the electron density $n = N/V$,
we thus arrive at the following relation:

$$\begin{aligned}
    \boxed{
        k_F^3
        = 3 \pi^2 n
    }
\end{aligned}$$

This result also justifies our assumption that $T = 0$:
we can accurately calculate the density $n$ for many conducting materials,
and this relation then gives $k_F$ and $\epsilon_F$.
It turns out that $\epsilon_F$ is usually very large
compared to the thermal energy $k_B T$ at reasonable temperatures,
so we can conclude that thermal fluctuations are negligible.

Now, $\epsilon_F$ is the highest single-electron energy,
but about the total $N$-particle energy $E^{(0)}$?

$$\begin{aligned}
    E^{(0)}
    = \matrixel{\mathrm{FS}}{\hat{H}_0}{\mathrm{FS}}
    = \sum_{s} \sum_{\vb{k}} \frac{\hbar^2 |\vb{k}|^2}{2 m} \matrixel{\mathrm{FS}}{\hat{c}_{s,\vb{k}}^\dagger \hat{c}_{s,\vb{k}}}{\mathrm{FS}}
\end{aligned}$$

Once again, we turn the sum over $\vb{k}$ into an integral,
and recognize the spin's irrelevance:

$$\begin{aligned}
    E^{(0)}
    &= \sum_{s} \frac{V}{(2 \pi)^3} \int_{-\infty}^\infty \frac{\hbar^2 |\vb{k}|^2}{2 m}
    \matrixel{\mathrm{FS}}{\hat{c}_{\vb{k}}^\dagger \hat{c}_{\vb{k}}}{\mathrm{FS}} \dd{\vb{k}}
    \\
    &= \frac{\hbar^2 V}{8 \pi^3 m} \int_{-\infty}^\infty |\vb{k}|^2 \: \Theta(k_F - |\vb{k}|) \dd{\vb{k}}
\end{aligned}$$

In spherical coordinates,
we evaluate the integral and find that $E^{(0)}$ is proportional to $k_F^5$:

$$\begin{aligned}
    E^{(0)}
    &= \frac{\hbar^2 V}{8 \pi^3 m} \int_0^{2 \pi}
    \int_0^\pi \int_0^\infty \Big( |\vb{k}|^2 \: \Theta(k_F - |\vb{k}|) \Big) |\vb{k}|^2 \sin(\theta) \dd{|\vb{k}|} \dd{\theta} \dd{\varphi}
    \\
    &= \frac{\hbar^2 V}{8 \pi^3 m} 4 \pi \int_0^{k_F} |\vb{k}|^4 \dd{|\vb{k}|}
    = \frac{\hbar^2 V}{2 \pi^2 m} \bigg[ \frac{|\vb{k}|^5}{5} \bigg]_0^{k_F}
    = \frac{\hbar^2 V}{10 \pi^2 m} k_F^5
\end{aligned}$$

In general, it is more useful to consider
the average kinetic energy per electron $E^{(0)} / N$,
which we find to be as follows, using that $k_F^3 = 3 \pi^2 n$:

$$\begin{aligned}
    \boxed{
        \frac{E^{(0)}}{N}
        = \frac{3 \hbar^2}{10 m} k_F^2
        = \frac{3}{5} \epsilon_F
    }
    \:\sim\: n^{2/3}
\end{aligned}$$

Traditionally, this is expressed using a dimensionless parameter $r_s$,
defined as the radius of a sphere containing a single electron,
measured in Bohr radii $a_0 \equiv 4 \pi \varepsilon_0 \hbar^2 / (e^2 m)$:

$$\begin{aligned}
        \frac{4 \pi}{3} (a_0 r_s)^3
        = \frac{1}{n}
        = \frac{3 \pi^2}{k_F^3}
        \quad \implies \quad
        r_s
        = \Big( \frac{3}{4 \pi a_0^3 n} \Big)^{1/3}
        = \Big( \frac{9 \pi}{4} \Big)^{1/3} \frac{1}{a_0 k_F}
\end{aligned}$$

Such that the ground state energy can be rewritten in Rydberg units of energy like so:

$$\begin{aligned}
    \frac{E^{(0)}}{N}
    = \frac{3 \hbar^2}{10 m} \frac{4 \pi \varepsilon_0 e^2}{4 \pi \varepsilon_0 e^2} \frac{a_0^2 k_F^2}{a_0^2}
    = \frac{3 e^2}{40 \pi \varepsilon_0} \Big( \frac{9 \pi}{4} \Big)^{2/3} \frac{1}{a_0 r_s^2}
    \approx \frac{2.21}{r_s^2} \; \mathrm{Ry}
\end{aligned}$$


## With interactions

To include Coulomb interactions, let us try
[time-independent pertubation theory](/know/concept/time-independent-perturbation-theory/).
Clearly, this will give better results when the interaction is relatively weak, if ever.

The Coulomb potential is proportional to the inverse distance,
and the average electron spacing is roughly $n^{-1/3}$,
so the interaction energy $E_\mathrm{int}$ should scale as $n^{1/3}$.
We already know that the kinetic energy $E_\mathrm{kin} = E^{(0)}$ scales as $n^{2/3}$,
meaning perturbation theory should be reasonable
if $1 \gg E_\mathrm{int} / E_\mathrm{kin} \sim n^{-1/3}$,
so in the limit of high density $n \to \infty$.

The two-body Coulomb interaction operator $\hat{W}$
is as follows in second-quantized form:

$$\begin{aligned}
    \hat{W}
    = \frac{1}{2 V} \sum_{s_1 s_2} \sum_{\vb{k}_1 \vb{k}_2} \sum_{\vb{q} \neq 0} \frac{e^2}{\varepsilon_0 |\vb{q}|^2}
    \hat{c}_{s_1, \vb{k}_1 + \vb{q}}^\dagger \hat{c}_{s_2, \vb{k}_2 - \vb{q}}^\dagger \hat{c}_{s_2, \vb{k}_2} \hat{c}_{s_1, \vb{k}_1}
\end{aligned}$$

The first-order correction $E^{(1)}$ to the ground state (i.e. Fermi sea) energy
is then given by:

$$\begin{aligned}
    E^{(1)}
    = \matrixel{\mathrm{FS}}{\hat{W}}{\mathrm{FS}}
    = \frac{e^2}{2 \varepsilon_0 V} \sum_{s_1 s_2} \sum_{\vb{k}_1 \vb{k}_2} \sum_{\vb{q} \neq 0} \frac{1}{|\vb{q}|^2}
    \matrixel{\mathrm{FS}}{
        \hat{c}_{s_1, \vb{k}_1 + \vb{q}}^\dagger \hat{c}_{s_2, \vb{k}_2 - \vb{q}}^\dagger \hat{c}_{s_2, \vb{k}_2} \hat{c}_{s_1, \vb{k}_1}
    }{\mathrm{FS}}
\end{aligned}$$

This inner product can only be nonzero
if the two creation operators $\hat{c}^\dagger$
are for the same orbitals as the two annihilation operators $\hat{c}$.
Since $\vb{q} \neq 0$, this means that $s_1 = s_2$,
and that momentum is conserved: $\vb{k}_2 = \vb{k}_1 \!+\! \vb{q}$.
And of course both $\vb{k}_1$ and $\vb{k}_1 \!+\! \vb{q}$
must be inside the Fermi sphere,
to avoid annihilating an empty orbital.
Let $s = s_1$ and $\vb{k} = \vb{k}_1$:

$$\begin{aligned}
    E^{(1)}
    &= \frac{e^2}{2 \varepsilon_0 V} \sum_{s} \sum_{\vb{k}} \sum_{\vb{q} \neq 0} \frac{1}{|\vb{q}|^2}
    \matrixel{\mathrm{FS}}{
        \hat{c}_{s, \vb{k} + \vb{q}}^\dagger \hat{c}_{s, \vb{k}}^\dagger \hat{c}_{s, \vb{k} + \vb{q}} \hat{c}_{s, \vb{k}}
    }{\mathrm{FS}}
    \\
    &= \frac{- e^2}{2 \varepsilon_0 V} \sum_{s} \sum_{\vb{k}} \sum_{\vb{q} \neq 0} \frac{1}{|\vb{q}|^2}
    \matrixel{\mathrm{FS}}{
        \big( \hat{c}_{s, \vb{k} + \vb{q}}^\dagger \hat{c}_{s, \vb{k} + \vb{q}}\big) \big(\hat{c}_{s, \vb{k}}^\dagger \hat{c}_{s, \vb{k}}\big)
    }{\mathrm{FS}}
    \\
    &= \frac{- e^2}{2 \varepsilon_0 V} \sum_{s} \sum_{\vb{k}} \sum_{\vb{q} \neq 0} \frac{1}{|\vb{q}|^2}
    \Theta(k_F - |\vb{k}|) \:\Theta(k_F - |\vb{k} \!+\! \vb{q}|)
\end{aligned}$$

Next, we convert the sum over $\vb{q}$ into an integral in spherical coordinates.
Clearly, $\vb{q}$ is the "jump" made by an electron from one orbital to another,
so the largest possible jump
goes from a point on the Fermi surface to the opposite point,
and thus has length $2 k_F$.
This yields the integration limit, and therefore leads to:

$$\begin{aligned}
    E^{(1)}
    &= \frac{- e^2}{(2 \pi)^3 \varepsilon_0} \sum_{\vb{k}}
    \int_0^{2 \pi} \!\!\int_0^\pi \!\!\int_0^\infty \Theta(k_F \!-\! |\vb{k}|) \: \Theta(k_F \!-\! |\vb{k} \!+\! \vb{q}|) \frac{|\vb{q}|^2}{|\vb{q}|^2}
    \sin(\theta_q) \dd{|\vb{q}|} \dd{\theta_q} \dd{\varphi_q}
    \\
    &= \frac{- e^2}{2 \pi^2 \varepsilon_0} \sum_{\vb{k}}
    \int_0^{2 k_F} \Theta(k_F \!-\! |\vb{k}|) \: \Theta(k_F \!-\! |\vb{k} \!+\! \vb{q}|) \dd{|\vb{q}|}
\end{aligned}$$

Where we have used that the direction of $\vb{q}$,
i.e. $(\theta_q,\varphi_q)$, is irrelevant,
as long as we define $\theta_k$ as
the angle between $\vb{q}$ and $\vb{k} \!+\! \vb{q}$
when we go to spherical coordinates $(|\vb{k}|, \theta_k, \varphi_k)$ for $\vb{k}$:

$$\begin{aligned}
    E^{(1)}
    &= \frac{- e^2 V}{16 \pi^5 \varepsilon_0} \int_0^{2 k_F} \!\!\!\!\int_0^{2 \pi} \!\!\!\int_0^\pi \!\!\!\int_0^\infty
    \!\Theta(k_F \!-\! |\vb{k}|) \: \Theta(k_F \!-\! |\vb{k} \!+\! \vb{q}|)
    \: |\vb{k}|^2 \sin(\theta_k) \dd{|\vb{k}|} \dd{\theta_k} \dd{\varphi_k} \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{16 \pi^5 \varepsilon_0} \int_0^{2 k_F} \!\!\!\!\int_0^{2 \pi} \!\!\!\int_0^\pi \!\!\!\int_0^{k_F}
    \!\Theta(k_F \!-\! |\vb{k} \!+\! \vb{q}|)
    \: |\vb{k}|^2 \sin(\theta_k) \dd{|\vb{k}|} \dd{\theta_k} \dd{\varphi_k} \dd{|\vb{q}|}
\end{aligned}$$

Unfortunately, this last step function is less easy to translate into integration limits.
In effect, we are trying to calculate the intersection volume of two spheres,
both with radius $k_F$, one centered on the origin (for $\vb{k}$),
and the other centered on $\vb{q}$ (for $\vb{k} \!+\! \vb{q}$).
Imagine a triangle with side lengths $|\vb{k}|$, $|\vb{q}|$ and $|\vb{k} \!+\! \vb{q}|^2$,
where $\theta_k$ is the angle between $|\vb{k}|$ and $|\vb{k} \!+\! \vb{q}|$.
The *law of cosines* then gives the following relation:

$$\begin{aligned}
    |\vb{k}|^2
    = |\vb{q}|^2 + |\vb{k} \!+\! \vb{q}|^2 - 2 |\vb{q}| |\vb{k} \!+\! \vb{q}| \cos(\theta_k)
\end{aligned}$$

We already know that $|\vb{k}| < k_F$ and $0 < |\vb{q}| < 2 k_F$,
so by isolating for $\cos(\theta_k)$,
we can obtain bounds on $\theta_k$ and $|\vb{k}|$.
Let $|\vb{k}| \to k_F$ in both cases, then:

$$\begin{aligned}
    \cos(\theta_k)
    = \frac{|\vb{k} \!+\! \vb{q}|^2 + |\vb{q}|^2 - |\vb{k}|^2}{2 |\vb{k} \!+\! \vb{q}| |\vb{q}|}
    &\:\:\underset{|\vb{q}| \to 0}{>}\:\:\: \frac{k_F^2 + |\vb{q}|^2 - k_F^2}{2 k_F |\vb{q}|}
    = \frac{|\vb{q}|}{2 k_F}
    \\
    &\underset{|\vb{q}| \to 2 k_F}{<}\:\: \frac{k_F^2 + 4 k_F ^2 - k_F^2}{2 k_F 2 k_F}
    = 1
\end{aligned}$$

Meaning that $0 < \theta_k < \arccos{|\vb{q}| / (2 k_F)}$.
To get a lower limit for $|\vb{k}|$, we "cheat" by artificially demanding
that $\vb{k}$ does not cross the halfway point between the spheres,
with the result that $|\vb{k}| \cos(\theta_k) > |\vb{q}|/2$.
Then, thanks to symmetry (both spheres have the same radius),
we just multiply the integral by $2$,
for $\vb{k}$ on the other side of the halfway point.

Armed with these integration limits, we return to calculating $E^{(1)}$,
substituting $\xi \equiv \cos(\theta_k)$:

$$\begin{aligned}
    E^{(1)}
    &= \frac{- e^2 V}{16 \pi^5 \varepsilon_0} 2 \int_0^{2 k_F} \!\!\!\int_0^{2 \pi} \!\!\int_0^{\arccos{|\vb{q}| / (2 k_F)}}
    \!\!\int_{|\vb{q}|/(2 \cos{\theta_k})}^{k_F} |\vb{k}|^2 \sin(\theta_k) \dd{|\vb{k}|} \dd{\theta_k} \dd{\varphi_k} \dd{|\vb{q}|}
    \\
    &= \frac{e^2 V}{8 \pi^5 \varepsilon_0} 2 \pi \int_0^{2 k_F} \!\!\!\int_1^{|\vb{q}| / (2 k_F)}
    \!\!\int_{|\vb{q}|/(2 \xi)}^{k_F} |\vb{k}|^2 \frac{\sin(\theta_k)}{\sin(\theta_k)} \dd{|\vb{k}|} \dd{\xi} \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \int_0^{2 k_F} \!\!\!\int_{|\vb{q}| / (2 k_F)}^1
    \!\!\int_{|\vb{q}|/(2 \xi)}^{k_F} |\vb{k}|^2 \dd{|\vb{k}|} \dd{\xi} \dd{|\vb{q}|}
\end{aligned}$$

Where we have used that $\varphi_k$ does not appear in the integrand.
Evaluating these integrals:

$$\begin{aligned}
    E^{(1)}
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \int_0^{2 k_F} \!\!\!\int_{|\vb{q}| / (2 k_F)}^1
    \bigg[ \frac{|\vb{k}|^3}{3} \bigg]_{|\vb{q}|/(2 \xi)}^{k_F} \dd{\xi} \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \int_0^{2 k_F} \!\!\!\int_{|\vb{q}| / (2 k_F)}^1
    \bigg( \frac{k_F^3}{3} - \frac{|\vb{q}|^3}{24 \xi^3} \bigg) \dd{\xi} \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \int_0^{2 k_F}
    \bigg[ \frac{k_F^3}{3} x + \frac{|\vb{q}|^3}{48 \xi^2} \bigg]_{|\vb{q}| / (2 k_F)}^1 \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \int_0^{2 k_F}
    \bigg( \frac{k_F^3}{3} + \frac{|\vb{q}|^3}{48} - \frac{k_F^2 |\vb{q}|}{4} \bigg) \dd{|\vb{q}|}
    \\
    &= \frac{- e^2 V}{4 \pi^4 \varepsilon_0} \bigg[ \frac{k_F^3 |\vb{q}|}{3} + \frac{|\vb{q}|^4}{192} - \frac{k_F^2 |\vb{q}|^2}{8} \bigg]_0^{2 k_F}
    \\
    &= \frac{- e^2 V}{16 \pi^4 \varepsilon_0} k_F^4
    = \frac{- e^2 N}{16 \pi^4 \varepsilon_0 n} k_F^4
    = -\frac{3 e^2 N}{16 \pi^2 \varepsilon_0} k_F
\end{aligned}$$

Per particle, the first-order energy correction $E^{(1)}$
is therefore found to be as follows:

$$\begin{aligned}
    \boxed{
        \frac{E^{(1)}}{N}
        = -\frac{3 e^2}{16 \pi^2 \varepsilon_0} k_F
    }
\end{aligned}$$

This can also be written using the parameter $r_s$ introduced above, leading to:

$$\begin{aligned}
    \frac{E^{(1)}}{N}
    = -\frac{3 e^2}{16 \pi^2 \varepsilon_0} \frac{a_0 k_F}{a_0}
    = -\frac{3 e^2}{16 \pi^2 \varepsilon_0} \Big( \frac{9 \pi}{4} \Big)^{1/3} \frac{1}{a_0 r_s}
\end{aligned}$$

Consequently, for sufficiently high densities $n$,
the total energy $E$ per particle is given by:

$$\begin{aligned}
    \boxed{
        \frac{E}{N}
        \approx \bigg( \frac{2.21}{r_s^2} - \frac{0.92}{r_s} \bigg) \; \mathrm{Ry}
    }
\end{aligned}$$

Unfortunately, this is as far as we can go.
In theory, the second-order energy correction $E^{(2)}$ is as shown below,
but it turns out that it (and all higher orders) diverge:

$$\begin{aligned}
    E^{(2)}
    = \sum_{\Psi_n \neq \mathrm{FS}} \frac{\big| \matrixel{\mathrm{FS}}{\hat{W}}{\Psi_n} \big|^2}{E^{(0)} - E_n}
\end{aligned}$$

The only cure for this is to go to infinite order,
where all the infinities add up to a finite result.



## References
1.  H. Bruus, K. Flensberg,
    *Many-body quantum theory in condensed matter physics*,
    2016, Oxford.