Ian Jauslin
summaryrefslogtreecommitdiff
blob: e2a31b0134ee3cbba28a768e1b6d6c6693bba70b (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
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
1123
1124
1125
1126
1127
1128
1129
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
1161
1162
1163
1164
1165
1166
1167
1168
1169
1170
1171
1172
1173
1174
1175
1176
1177
1178
1179
1180
1181
1182
1183
1184
1185
1186
1187
1188
1189
1190
1191
1192
1193
1194
1195
1196
1197
1198
1199
1200
1201
1202
1203
1204
1205
1206
1207
1208
1209
1210
1211
1212
1213
1214
1215
1216
1217
1218
1219
1220
1221
1222
1223
1224
1225
1226
1227
1228
1229
1230
1231
1232
1233
1234
1235
1236
1237
1238
1239
1240
1241
1242
1243
1244
1245
1246
1247
1248
1249
1250
1251
1252
1253
1254
1255
1256
1257
1258
1259
1260
1261
1262
1263
1264
1265
1266
1267
1268
1269
1270
1271
1272
1273
1274
1275
1276
1277
1278
1279
1280
1281
1282
1283
1284
1285
1286
1287
1288
1289
1290
1291
1292
1293
1294
1295
1296
1297
1298
1299
1300
1301
1302
1303
1304
1305
1306
1307
1308
1309
1310
1311
1312
1313
1314
1315
1316
1317
1318
1319
1320
1321
1322
1323
1324
1325
1326
1327
1328
1329
1330
1331
1332
1333
1334
1335
1336
1337
1338
1339
1340
1341
1342
1343
1344
1345
1346
1347
1348
1349
1350
1351
1352
1353
1354
1355
1356
1357
1358
1359
1360
1361
1362
1363
1364
1365
1366
1367
1368
1369
1370
1371
1372
1373
1374
1375
1376
1377
1378
1379
1380
1381
1382
1383
1384
1385
1386
1387
1388
1389
1390
1391
1392
1393
1394
1395
1396
1397
1398
1399
1400
1401
1402
1403
1404
1405
1406
1407
1408
1409
1410
1411
1412
1413
1414
1415
1416
1417
1418
1419
1420
1421
1422
1423
1424
1425
1426
1427
1428
1429
1430
1431
1432
1433
1434
1435
1436
1437
1438
1439
1440
1441
1442
1443
1444
1445
1446
1447
1448
1449
1450
1451
1452
1453
1454
1455
1456
1457
1458
1459
1460
1461
1462
1463
1464
1465
1466
1467
1468
1469
1470
1471
1472
1473
1474
1475
1476
1477
1478
1479
1480
1481
1482
1483
1484
1485
1486
1487
1488
1489
1490
1491
1492
1493
1494
1495
1496
1497
1498
1499
1500
1501
1502
1503
1504
1505
1506
1507
1508
1509
1510
1511
1512
1513
1514
1515
1516
1517
1518
1519
1520
1521
1522
1523
1524
1525
1526
1527
1528
1529
1530
1531
1532
1533
1534
1535
1536
1537
1538
1539
1540
1541
1542
1543
1544
1545
1546
1547
1548
1549
1550
1551
1552
1553
1554
1555
1556
1557
1558
1559
1560
1561
1562
1563
1564
1565
1566
1567
1568
1569
1570
1571
1572
1573
1574
1575
1576
1577
1578
1579
1580
1581
1582
1583
1584
1585
1586
1587
1588
1589
1590
1591
1592
1593
1594
1595
1596
1597
1598
1599
1600
1601
1602
1603
1604
1605
1606
1607
1608
1609
1610
1611
1612
1613
1614
1615
1616
1617
1618
1619
1620
1621
1622
1623
1624
1625
1626
1627
1628
1629
1630
1631
1632
1633
1634
1635
1636
1637
1638
1639
1640
1641
1642
1643
1644
1645
1646
1647
1648
1649
1650
1651
1652
1653
1654
1655
1656
1657
1658
1659
1660
1661
1662
1663
1664
1665
1666
1667
1668
1669
1670
1671
1672
1673
1674
1675
1676
1677
1678
1679
1680
1681
1682
1683
1684
1685
1686
1687
1688
1689
1690
1691
1692
1693
1694
1695
1696
1697
1698
1699
1700
1701
1702
1703
1704
1705
1706
1707
1708
1709
1710
\documentclass{ian}

\usepackage{dsfont}
\usepackage{iantheo}
\usepackage{largearray}
\usepackage{point}
\usepackage{etoolbox}

\begin{document}

\pagestyle{empty}

\hbox{}
\hfil{\bf\LARGE
The Simplified approach to the Bose gas\par
\medskip
\hfil without translation invariance
}
\vfill

\hfil{\bf\large Ian Jauslin}\par
\hfil{\it Department of Mathematics, Rutgers University}\par
\hfil{\tt\color{blue}\href{mailto:ian.jauslin@rutgers.edu}{ian.jauslin@rutgers.edu}}\par
\vskip20pt

\vfill


\hfil {\bf Abstract}\par
\smallskip

The Simplified approach to the Bose gas was introduced by Lieb in 1963 to study the ground state of systems of interacting Bosons.
In a series of recent papers, it has been shown that the Simplified approach exceeds earlier expectations, and gives asymptotically accurate predictions at both low and high density.
In the intermediate density regime, the qualitative predictions of the Simplified approach have also been found to agree very well with Quantum Monte Carlo computations.
Until now, the Simplified approach had only been formulated for translation invariant systems, thus excluding external potentials, and non-periodic boundary conditions.
In this paper, we extend the formulation of the Simplified approach to a wide class of systems without translation invariance.
This also allows us to study observables in translation invariant systems whose computation requires the symmetry to be broken.
Such an observable is the momentum distribution, which counts the number of particles in excited states of the Laplacian.
In this paper, we show how to compute the momentum distribution in the Simplified approach, and show that, for the Simple Equation, our prediction matches up with Bogolyubov's prediction at low densities, for momenta extending up to the inverse healing length.

\vfill

\tableofcontents

\eject

\setcounter{page}1
\pagestyle{plain}

\section{Introduction}
\indent
The Bose gas is one of the simplest models in quantum statistical mechanics, and yet it has a rich and complex phenomenology.
As such, it has garnered much attention from the mathematical physics community for over half a century.
It consists in infinitely many identical Bosons and is used to model a wide range of physical systems, from photons in black body radiation to gasses of helium atoms.
Whereas photons do not directly interact with each other, helium atoms do, and such an interaction makes studying such systems very challenging.
To account for interactions between Bosons, Bogolyubov\-~\cite{Bo47} introduced a widely used approximation scheme that accurately predicts many observables\-~\cite{LHY57} {\it in the low density} regime.
Even though Bogolyubov theory is not mathematically rigorous, it has allowed mathematical physicists to develop the necessary intuition to prove a wide variety of results about the Bose gas, such as the low density expansion of the ground state energy of the Bose gas in the thermodynamic limit\-~\cite{Dy57,LY98,YY09,FS20,BCS21,FS22}, as well as many other results in scaling limits other than the thermodynamic limit (see\-~\cite{Sc22} for a review, as well as, among many others, \cite{LSY00,LS02,NRS16,BBe18,BBe19,DSY19,BBe20,DS20,NT21,BSS22,BSS22b,HST22,NNe22}).
In this note, we will focus on the ground state in the thermodynamic limit.
\bigskip

\indent
In 1963, E.H.\-~Lieb\-~\cite{Li63,LS64,LL64} introduced a new approximation scheme to compute properties of the ground state of Bose gasses, called the {\it Simplified approach}, which has recently been found to yield surprisingly accurate results\-~\cite{CJL20,CJL21,CHe21,Ja22}.
Indeed, while Bogolyubov theory is accurate at low densities, the Simplified approach has been shown to yield asymptotically accurate results at both {\it low and high} densities\-~\cite{CJL20,CJL21} for interaction potentials that are of positive type, as well as reproduce the qualitative behavior of the Bose gas at intermediate densities\-~\cite{CHe21}.
In addition to providing a promising tool to study the Bose gas, the derivation of the Simplified approach is different enough from Bogolyubov theory that it may give novel insights into longstanding open problems about the Bose gas.
\bigskip

\indent
The original derivation of the Simplified approach\-~\cite{Li63} is quite general, and applies to any translation invariant system (it even works for Coulomb\-~\cite{LS64} and hard-core\-~\cite{CHe21} interactions).
In the present paper, we extend this derivation to systems that break translation invariance.
This allows us to formulate the Simplified approach for systems with external potentials, and with a large class of boundary conditions.
In addition, it allows us to compute observables in systems with translation invariance, but whose computation requires breaking the translation invariance.
We will discuss an example of such an observable: the momentum distribution.
\bigskip

\indent
The momentum distribution $\mathcal M(k)$ is the probability of finding a particle in the state $e^{ikx}$.
Bose gasses are widely expected to form a Bose-Einstein condensate, although this has still not been proven (at least for continuum interacting gasses in the thermodynamic limit).
From a mathematical point of view, Bose-Einstein condensation is defined as follows: if the Bose gas consists of $N$ particles, the average number of particles in the constant state (corresponding to $k=0$ in $e^{ikx}$) is of order $N$.
The {\it condensate fraction} is defined as the proportion of particles in the constant state.
The momentum distribution is an extension of the condensate fraction to a more general family of states.
In particular, computing $\mathcal M(k)$ for $k\neq 0$ amounts to counting particles that are {\it not} in the condensate.
This quantity has been used in the recent proof\-~\cite{FS20,FS22} of the energy asymptotics of the Bose gas at low density.
A numerical computation of the prediction of the Simplified approach for $\mathcal M(k)$ have been published in\-~\cite{Ja23b}.
\bigskip

\indent
The main results in this paper fall into two categories.
First, we will derive the Simplified approach without assuming translation invariance, see Theorem\-~\ref{theo:simple}.
To do so, we will make the so-called ``factorization assumption'', on the marginals of the ground state wavefunction, see Assumption\-~\ref{assum:factorization}.
This allows us to derive a Simplified approach for a wide variety of situations in which translation symmetry breaking is violated, such as in the presence of external potentials.
Second, we compute a prediction for the momentum distribution using the Simplified approach.
The Simplified approach does not allow us to compute the ground state wavefunction directly, so to compute observables, such as the momentum distribution, we use the Hellmann-Feynman technique and add an operator to the Hamiltonian.
In the case of the momentum distribution, this extra operator is a projector onto $e^{ikx}$, which breaks the translation invariance of the system.
In Theorem\-~\ref{theo:Nk}, we show how to compute the momentum distribution in the Simplified approach using the general result of Theorem\-~\ref{theo:simple}.
In addition, we check that the prediction is credible, by comparing it to the prediction of Bogolyubov theory, and find that both approaches agree at low densities and small $k$, see Theorem\-~\ref{theo:Nk_bog}.
\bigskip

\indent
The rest of the paper is structured as follows.
In Section\-~\ref{sec:model}, we specify the model and state the main results precisely.
We then prove Theorem\-~\ref{theo:simple} in Section\-~\ref{sec:simple}, Theorem\-~\ref{theo:Nk} in Section\-~\ref{sec:Nk_proof}, and Theorem\-~\ref{theo:Nk_bog} in Section\-~\ref{sec:Nk_bog}.
The proofs are largely independent and can be read in any order.
\bigskip

\section{The model and main results}\label{sec:model}
\indent
Consider $N$ Bosons in a box of volume $V$ denoted by $\Omega_V:=[-V^{\frac13}/2,V^{\frac13}/2]^3$, interacting with each other via a pair potential $v\in L_{1}(\Omega_V^2)$ that is symmetric under exchanges of particles: $v(x,y)\equiv v(y,x)$.
The Hamiltonian acts on $L_{2,\mathrm{sym}}(\Omega_V^N)$ as
\begin{equation}
  \mathcal H:=
  -\frac12\sum_{i=1}^N\Delta_i
  +
  \sum_{1\leqslant i<j\leqslant N}v(x_i,x_j)
  +
  \sum_{i=1}^N P_i
  \label{ham}
\end{equation}
where $\Delta_i\equiv\partial_{x_i}^2$ is the Laplacian with respect to the position of the $i$-th particle and $P_i$ is an extra single-particle term of the following form: given a self-adjoint operator $\varpi$ on $L_2(\Omega_V)$,
\begin{equation}
  P_i:=\mathds 1^{\otimes i-1}\otimes \varpi\otimes\mathds 1^{\otimes N-i}
  .
  \label{Ppi}
\end{equation}
For instance, if we take $\varpi$ to be a multiplication operator by a function $v_0$, then $\sum_i P_i$ is the contribution of the external potential $v_0$.
Or $\varpi$ could be a projector onto $e^{ikx}$, which is what we will do below to compute the momentum distribution.
Because $P_i$ acts on a single particle, it breaks translational symmetry as soon as it is not constant.
\bigskip

\indent
We may impose any boundary condition on the box, as long as the Laplacian is self-adjoint.
We will consider the thermodynamic limit, in which $N,V\to\infty$, such that
\begin{equation}
  \frac NV=\rho
\end{equation}
is fixed.
We consider the ground state $\psi_0$, which is the eigenfunction of $\mathcal H$ with the lowest eigenvalue $E_0$:
\begin{equation}
  \mathcal H\psi_0=E_0\psi_0
  .
  \label{eigval}
\end{equation}
(It is a standard argument to prove that $\psi_0$ exists, and is both real and non-negative.)
\bigskip

\indent
In order to take the thermodynamic limit, we will assume that $v$ is uniformly integrable in $V$:
\begin{equation}
  |v(x,y)|\leqslant \bar v(x,y)
  ,\quad
  \int_{\mathbb R^3} dy\ \bar v(x,y)\leqslant c
  \label{intv}
\end{equation}
where $\bar v$ and $c$ are independent of $V$.
In addition, we assume that, for any $f$ that is uniformly integrable in $V$,
\begin{equation}
  \int dx\ \varpi f(x)\leqslant c
  .
  \label{bound_varpi}
\end{equation}
\bigskip

\subsection{The Simplified approach without translation invariance}\label{sec:general}
\indent
The crucial idea of Lieb's construction\-~\cite{Li63} is to consider the wave function $\psi$ as a probability distribution, instead of the usual $|\psi|^2$.
Since $\psi\geqslant 0$, this can be done by normalizing $\psi$ by its $L_1$ norm.
We then define the $i$-th marginal of $\psi$ as
\begin{equation}
  g_i(x_1,\cdots,x_i)
  :=
  \frac{\int\frac{dx_{i+1}}V\cdots\frac{dx_N}V\ \psi(x_1,\cdots,x_N)}{\int\frac{dy_{1}}V\cdots\frac{dy_N}V\ \psi(y_1,\cdots,y_N)}
  \equiv
  V^i\frac{\int dx_{i+1}\cdots dx_N\ \psi(x_1,\cdots,x_N)}{\int dy_{1}\cdots dy_N\ \psi(y_1,\cdots,y_N)}
  .
  \label{gdef}
\end{equation}
In particular, for $i\in\{2,\cdots,N\}$,
\begin{equation}
  \int\frac{dx_i}V\ g_i(x_1,\cdots,x_i)=g_{i-1}(x_1,\cdots,x_{i-1})
  ,\quad
  \int\frac{dx}V\ g_1(x)=1
  .
  \label{grec}
\end{equation}
Because of the symmetry of $\psi$ under exchanges of particles, $g_i$ is symmetric under $x_i\leftrightarrow x_j$.
\bigskip

\indent
Inspired by\-~\cite{Li63}, we will make the following approximation.
\bigskip

\theoname{Assumption}{Factorization}\label{assum:factorization}
  For $i=2,3,4$,
  \begin{equation}
    g_i(x_1,\cdots,x_i)
    =
    \prod_{1\leqslant j<l\leqslant i}
    W_i(x_j,x_l)
    \label{g_factorized}
  \end{equation}
  with
  \begin{equation}
    W_i(x,y)=f_i(x)f_i(y)(1-u_i(x,y))
    \label{W_fact}
  \end{equation}
  in which $f_i$ and $u_i$ are bounded independently of $V$ and $u_i$ is uniformly integrable in $V$:
  \begin{equation}
    |u_i(x,y)|\leqslant\bar u_i(x,y)
    ,\quad
    \int dy\ \bar u_i(x,y)\leqslant c_i
    \label{assum_bound}
  \end{equation}
  with $c_i$ independent of $V$.
  We further assume that, for $i=1,2,3$,
  \begin{equation}
    \lim_{V\to\infty}\int dx_i\ \Delta_{x_i} g_i(x_1,\cdots,x_i)=0
  \end{equation}
  in other words, these boundary terms vanish in the thermodynamic limit.
\endtheo
\bigskip

In other words, $g_i$ factorizes exactly as a product of pair terms $W_i$.
The $f_i$ in $W_i$ allow for $W_i$ to be modulated by a slowly varying density, which is the main novelty of this paper compared to\-~\cite{Li63}.
The inequality\-~(\ref{assum_bound}) ensures that $u_i$ decays sufficiently fast on the microscopic scale.
Note that, by the symmetry under exchanges of particles, $u_i(x,y)\equiv u_i(y,x)$.
\bigskip

\indent
Here, we use the term ``assumption'' because it leads to the Simplified approach.
However, it is really an {\it approximation} rather than an assumption: this factorization will certainly not hold true exactly.
At best, one might expect that the assumption holds approximately in the limit of small and large $\rho$, and for distant points, as numerical evidence suggests in the translation invariant case.
In the present paper, we will not attempt a proof that this approximation is accurate, and instead explore its consequences.
Suffice it to say that this approximation is one of {\it statistical independence} that is reminiscent of phenomena arising in statistical mechanics when the density is low, that is, when the interparticle distances are large.
In the current state of the art, we do not have much in the way of an explanation for why this statistical independence should hold; instead, we have extensive evidence, both numerical\-~\cite{CHe21} and analytical\-~\cite{CJL20,CJL21}, that this approximation leads to very accurate predictions.
\bigskip

\indent
The equations of the Simplified approach are derived from Assumption\-~\ref{assum:factorization}, using the eigenvalue equation\-~(\ref{eigval}) along with
\begin{equation}
  \int\frac{dx}V\ g_1(x)=1
  \label{g11}
\end{equation}
\begin{equation}
  \int\frac{dy}V\ g_2(x,y)=g_1(x)
  \label{g2g1}
\end{equation}
\begin{equation}
  \int\frac{dz}V\ g_3(x,y,z)=g_2(x,y)
  \label{g3g2}
\end{equation}
\begin{equation}
  \int\frac{dz}V\frac{dt}V\ g_4(x,y,z,t)=g_2(x,y)
  \label{g4g2}
\end{equation}
(all of which follow from\-~(\ref{grec})) to compute $u_i$ and $f_i$.
\bigskip

\indent
In the translation invariant case, the factorization assumption leads to an equation for $g_2$ alone, as $g_1$ is constant.
When translation invariance is broken, $g_1$ is no longer constant, and the Simplified approach consists in two coupled equations for $g_1$ and $g_2$.
We formulate these in terms of $g_1$ and $u_2$, with
\begin{equation}
  g_2(x,y)=:g_1(x)g_1(y)(1-u_2(x,y))
  .
\end{equation}
\bigskip

\theo{Theorem}\label{theo:simple}
  If $g_i$ satisfies Assumption\-~\ref{assum:factorization}, the eigenvalue equation\-~(\ref{eigval}) and\-~(\ref{g11})-(\ref{g4g2}), then $g_1$ and $u_2$ satisfy the two coupled equations
  \begin{equation}
    \left(
      -\frac\Delta 2
      +\left(\varpi-\left<\varpi\right>\right)
      +2\left(\mathcal E(x)-\left<\mathcal E(y)\right>\right)
      +\frac12\left(\bar A(x)-\left<\bar A\right>-\bar C(x)\right)
    \right)g_1(x)
    +\Sigma_1(x)
    =0
    \label{compleq_g1}
  \end{equation}
  and
  \begin{equation}
    \begin{largearray}
      \left(-\frac12(\Delta_x+\Delta_y)+v(x,y)-2\rho \bar K(x,y)+\rho^2\bar L(x,y)+\bar R_2(x,y)\right)
      g_1(x)g_1(y)(1-u_2(x,y))
      +\\\hfill+
      \Sigma_2(x,y)=0
      \label{compleq_g2}
    \end{largearray}
  \end{equation}
  where
  \begin{equation}
    \left<f\right>:=\int\frac{dy}V\ g_1(y)f(y)
    ,\quad
    \left<\varpi\right>\equiv \int\frac{dy}V\ \varpi g_1(y)
    \label{avgdef}
  \end{equation}
  \begin{equation}
    \bar S(x,y):=v(x,y)(1-u_2(x,y))
    ,\quad
    f_1\bar\ast f_2(x,y):=\int dz\ g_1(z)f_1(x,z)f_2(z,y)
  \end{equation}
  \begin{equation}
    \mathcal E(x):=
    \frac\rho2\int dy\ g_1(y)\bar S(x,y)
    ,\quad
    \bar A(x):=
    \rho^2\bar S\bar\ast u_2\bar\ast u_2(x,x)
    \label{EA}
  \end{equation}
  \begin{equation}
    \bar C(x):=
    2\rho^2\int dz\ g_1(z)u_2\bar\ast\bar S(x,z)
    +2\rho\int dy\ \varpi_y(g_1(y)u_2(x,y))
    .
    \label{C}
  \end{equation}
  \begin{equation}
    \bar K(x,y)
    :=
    \bar S\bar\ast u_2(x,y)
  \end{equation}
  \begin{equation}
    \begin{largearray}
      \bar L(x,y)
      :=
      \bar S\bar\ast u_2\bar\ast u_2(x,y)
      -2u_2\bar\ast(u_2(u_2\bar\ast\bar S))(x,y)
      +\\\hfill+
      \frac12\int dzdt\ g_1(z)g_1(t)\bar S(z,t) u_2(x,z)u_2(x,t)u_2(y,z)u_2(y,t)
    \end{largearray}
  \end{equation}
  \begin{equation}
    \begin{array}{r@{\ }>\displaystyle l}
      \bar R_2(x,y)
      =&
      2\left(\mathcal E(x)+\mathcal E(y)-2\left<\mathcal E\right>\right)
      +\left(\varpi_x+\varpi_y-2\left<\varpi\right>\right)
      +\\[0.3cm]&+
      \frac12\left(\bar A(x)+\bar A(y)-2\left<\bar A\right>-\bar C(x)-\bar C(y)\right)
      +2\rho u_2\bar\ast\left(u_2(\mathcal E-\left<\mathcal E\right>)\right)
      +\\[0.3cm]&+
      \rho\int dz\ \varpi_z(g_1(z)u_2(x,z)u_2(y,z))
      -
      \rho u_2\bar\ast u_2\left<\varpi\right>
      \label{R}
    \end{array}
  \end{equation}
  in which $\varpi_x$ is the action of $\varpi$ on the $x$-variable, and similarly for $\varpi_y$
  and
  \begin{equation}
    \Sigma_i\mathop{\longrightarrow}_{V\to\infty}0
  \end{equation}
  pointwise.
  Furthermore, the prediction for the energy per particle is
  \begin{equation}
    e:=\left<\mathcal E\right>+\left<\varpi\right>+\Sigma_0
    \label{simplen}
  \end{equation}
  where $\Sigma_0\to0$ as $V\to\infty$.
\endtheo
\bigskip

This theorem is proved in Section\-~\ref{sec:simple}.
\bigskip

\indent
Let us compare this to the equation for $u$ in the Simplified approach in the translation invariant case\-~\cite[(5)]{CHe21}, \cite[(3.15)]{Ja22}:
\begin{equation}
  -\Delta u(x)
  =
  (1-u(x))\left(v(x)-2\rho K(x)+\rho^2 L(x)\right)
  \label{compleq}
\end{equation}
\begin{equation}
  K:=
  u\ast S
  ,\quad
  S(y):=(1-u(y))v(y)
  \label{K}
\end{equation}
\nopagebreakaftereq
\begin{equation}
  L:=
  u\ast u\ast S
  -2u\ast(u(u\ast S))
  +\frac12
  \int dydz\ u(y)u(z-x)u(z)u(y-x)S(z-y)
  .
  \label{L}
\end{equation}
We will prove that these follow from Theorem\-~\ref{theo:simple}:
\bigskip

\theoname{Corollary}{Translation invariant case}\label{cor:check}
  In the translation invariant case $v(x,y)\equiv v(x-y)$ and $\varpi=0$ with periodic boundary conditions, if\-~(\ref{compleq_g1})-(\ref{compleq_g1}) has a unique translation invariant solution, then (\ref{compleq_g2}) reduces to\-~(\ref{compleq}) in the thermodynamic limit.
\endtheo
\bigskip

\indent
The idea of the proof is quite straightforward.
Equation\-~(\ref{compleq_g2}) is very similar to\-~(\ref{compleq}), but for the addition of the extra term $\bar R_2$.
An inspection of\-~(\ref{R}) shows that the terms in $\bar R_2$ are mostly of the form $f-\left<f\right>$, which vanish in the translation invariant case, and terms involving $\varpi$, which is set to 0 in the translation invariant case.
The only remaining extra term is $\bar C(x)+\bar C(y)$, which we will show vanishes in the translation invariant case due to the identity\-~(\ref{g2g1}).
\bigskip

\indent
Theorem\-~\ref{theo:simple} is quite general, and can be used to study a trapped Bose gas, in which there is an external potential $v_0$.
In this case, $\varpi$ is a multiplication operator by $v_0$.
A natural approach is to scale $v_0$ with the volume: $v_0(x)=\bar v_0(V^{-1/3}x)$ in such a way that the size of the trap grows as $V\to\infty$, thus ensuring a finite local density in the thermodynamic limit.
Following the ideas of Gross and Pitaevskii\-~\cite{Gr61,Pi61}, we would then expect to find that\-~(\ref{compleq_g1}) and\-~(\ref{compleq_g2}) decouple, and that (\ref{compleq_g2}) reduces to the translation invariant equation\-~(\ref{compleq}), with a density that is modulated over the trap.
However, the presence of $\bar R_2$ in\-~(\ref{compleq_g2}) and $\bar C$ in\-~(\ref{compleq_g1}) breaks this picture.
Further investigation of this question is warranted.


\subsection{The momentum distribution}\label{sec:Nk}
\indent
The momentum distribution for the Bose gas is defined as
\begin{equation}
  \mathcal M^{(\mathrm{Exact})}(k):=\frac1N\sum_{i=1}^N\left<\psi_0\right|P_i\left|\psi_0\right>
  \label{Mdef}
\end{equation}
where
\begin{equation}
  \varpi f:=\epsilon| e^{ikx}\big>\big< e^{ikx}|f
  \equiv
  \epsilon e^{ikx}\int dy\ e^{-iky}f(y)
  \label{varpiNk}
\end{equation}
and $P_i$ is defined as in\-~(\ref{Ppi}):
\begin{equation}
  P_i\psi(x_1,\cdots,x_N)= \epsilon e^{ikx_i}\int dy_y\ e^{iky_i}\psi(x_1,\cdots,x_{i-1},y_i,x_{i+1},\cdots,x_N)
\end{equation}
Equivalently,
\begin{equation}
  \mathcal M^{(\mathrm{Exact})}(k)=\frac\partial{\partial\epsilon}\left. \frac{E_0}N\right|_{\epsilon=0}
\end{equation}
where $E_0$ is the energy in\-~(\ref{eigval}) for the Hamiltonian\-~(\ref{ham}).
Using the Simplified approach, we do not have access to the ground state wavefunction, so we cannot compute $\mathcal M$ using\-~(\ref{Mdef}).
Instead, we use the Hellmann-Feynman theorem, which consists in adding $\sum_iP_i$ to the Hamiltonian.
However, doing so breaks the translational symmetry.
This is why Theorem\-~\ref{theo:simple} is needed to compute the momentum distribution.
(A similar computation was done in\-~\cite{CHe21}, but, there, the derivation of the momentum distribution for the Simplified approach was taken for granted.)
\bigskip

\indent
By Theorem\-~\ref{theo:simple}, and, in particular, (\ref{simplen}), we obtain a natural definition of the prediction of the Simplified approach for the momentum distribution:
\begin{equation}
  \mathcal M(k):=\frac{\partial}{\partial\epsilon}\left.\left(\left<\mathcal E\right>+\left<\varpi\right>\right)\right|_{\epsilon=0}
  .
\end{equation}
\bigskip

\theoname{Theorem}{Momentum distribution}\label{theo:Nk}
  Under the assumptions of Theorem\-~\ref{theo:simple}, using periodic boundary conditions, if $v$ is translation invariant and $\varpi=0$, then, if $k\neq 0$, in the thermodynamic limit,
  \begin{equation}
    \mathcal M(k)=\frac{\partial}{\partial\epsilon}\left.\frac\rho2\int dx\ (1-u(x))v(x)\right|_{\epsilon=0}
  \end{equation}
  where
  \begin{equation}
    -\Delta u(x)=(1-u(x))v(x)-2\rho K(x)+\rho^2L(x)+\epsilon F(x)
  \end{equation}
  where $K$ and $L$ are those of the translation invariant Simplified approach\-~(\ref{K})-(\ref{L}) and
  \begin{equation}
    F(x):=-2\hat u(-k)\cos(kx)
    .
    \label{F}
  \end{equation}
\endtheo
\bigskip

\indent
We thus compute the momentum distribution.
To check that our prediction is plausible, we compare it to the Bogolyubov prediction, which can easily be derived from\-~\cite[Appendix\-~A]{LSe05}:
\begin{equation}
  \mathcal M^{(\mathrm{Bogolyubov})}(k)=-\frac1{2\rho}\left(1-\frac{k^2+2\rho\hat v(k)}{\sqrt{k^4+4k^2\rho\hat v(k)}}\right)
\end{equation}
(this can be obtained by differentiating\-~\cite[(A.26)]{LSe05} with respect to $\epsilon(k)$, which returns the number of particles in the state $e^{ikx}$, which we divide by $\rho$ to obtain the momentum distribution).
Actually, following the ideas of\-~\cite{LHY57}, we replace $\hat v$ by a so-called ``pseudopotential'', which consists in replacing $v$ by a Dirac delta function, while preserving the scattering length:
\begin{equation}
  \hat v(k)=4\pi a
\end{equation}
where the scattering length $a$ is defined in\-~\cite[Appendix\-~C]{LSe05}.
Thus,
\begin{equation}
  \mathcal M^{(\mathrm{Bogolyubov})}(k)=-\frac1{2\rho}\left(1-\frac{k^2+8\pi\rho a}{\sqrt{k^4+16\pi k^2\rho a}}\right)
  .
  \label{Mbog}
\end{equation}
\bigskip

\indent
We prove that, for the Simple Equation, as $\rho\to0$, the prediction for the momentum distribution coincides with Bogolyubov's, for $|k|\lesssim\sqrt{\rho a}$.
The length scale $1/\sqrt{\rho a}$ is called the {\it healing length}, and is the distance at which pairs of particles correlate\-~\cite{FS20}.
It is reasonable to expect the Bogolyubov approximation to break down beyond this length scale.
\bigskip

\indent
The momentum distribution for the Simple equation, following the prescription detailed in\-~\cite{CJL20,CJL21,CHe21,Ja22}, is defined as
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)=\frac{\partial}{\partial\epsilon}\left.\frac\rho2\int dx\ (1-u(x))v(x)\right|_{\epsilon=0}
  \label{M_simpleq}
\end{equation}
where\-~\cite[(1.1)-(1.2)]{CJL20}
\begin{equation}
  -\Delta u(x)=(1-u(x))v(x)-4eu+2\rho e u\ast u+\epsilon F(x)
  ,\quad
  e:=\frac\rho2\int dx\ (1-u(x))v(x)
  \label{simpleq}
\end{equation}
where $F$ was defined in\-~(\ref{F}).
\bigskip

\theo{Theorem}\label{theo:Nk_bog}
  Assume that $v$ is translation and rotation invariant ($v(x,y)\equiv v(|x-y|)$), and consider periodic boundary conditions.
  We rescale $k$:
  \begin{equation}
    \kappa:=\frac{k}{2\sqrt e}
  \end{equation}
  we have, for all $\kappa\in\mathbb R^3$,
  \begin{equation}
    \lim_{e\to0}\rho\mathcal M^{(\mathrm{simpleq})}(2\sqrt e\kappa)
    =\lim_{e\to0}\rho\mathcal M^{(\mathrm{Bogolyubov})}(2\sqrt e\kappa)
    =-\frac12\left(1-\frac{\kappa^2+1}{\sqrt{(\kappa^2+1)^2-1}}\right)
    .
    \label{Msimpleqbog}
  \end{equation}
\endtheo
\bigskip

\indent
The rotation invariance of $v$ is presumably not necessary.
However, the proof of this theorem is based on\-~\cite{CJL21}, where rotational symmetry was assumed for convenience.


\section{The Simplified approach without translation invariance, proof of Theorem \expandonce{\ref{theo:simple}}}\label{sec:simple}

\subsection{Factorization}
\indent
We will first compute $f_i$ and $u_i$ in Assumption\-~\ref{assum:factorization}.
\bigskip

\subsubsection{Factorization of $g_2$}
\indent
We start by considering $g_2$.
\bigskip

\theo{Lemma}\label{lemma:g2}
  Assumption\-~\ref{assum:factorization} with $i=2$ and\-~(\ref{g11})-(\ref{g2g1}) imply that
  \begin{equation}
    g_2(x,y)=g_1(x)g_1(y)(1-u(x,y))(1+O(V^{-2}))
    .
  \end{equation}
\endtheo

\indent\underline{Proof}:
  Assumption\-~\ref{assum:factorization} implies
  \begin{equation}
    g_2(x,y)=f_2(x)f_2(y)(1-u_2(x,y))
    .
  \end{equation}
  and by\-~(\ref{g2g1}),
  \begin{equation}
    g_1(x)=f_2(x)\int \frac{dy}V f_2(y)(1-u_2(x,y))
    .
    \label{g1_fact}
  \end{equation}
  \bigskip

  \point
  Let us first take an expansion to order $V^{-1}$.
  By~\-(\ref{assum_bound})
  \begin{equation}
    \int\frac{dy}V\ f_2(y)u_2(x,y)=O(V^{-1})
  \end{equation}
  and so
  \begin{equation}
    g_1(x)=f_2(x)\left(\int\frac{dy}V f_2(y)+O(V^{-1})\right)
    .
    \label{g1f2}
  \end{equation}
  Applying $\int\frac{dx}V\cdot$ to both sides of\-~(\ref{g1f2}), we find that
  \begin{equation}
    \int\frac{dy}Vf_2(y)=1+O(V^{-1})
  \end{equation}
  so\-~(\ref{g1f2}) yields
  \begin{equation}
    f_2(x)=g_1(x)(1+O(V^{-1}))
    .
    \label{f1V1}
  \end{equation}
  \bigskip

  \point
  We now push the expansion to order $V^{-2}$.
  Inserting\-~(\ref{f1V1}) into\-~(\ref{g1_fact}),
  \begin{equation}
    g_1(x)=f_2(x)\int\frac{dy}V\ f_2(y)-g_1(x)\left(\int\frac{dy}V\ g_1(y)u_2(x,y)+O(V^{-2})\right)
    .
  \end{equation}
  However, by\-~(\ref{g2g1}),
  \begin{equation}
    g_1(x)\int\frac{dy}V\ g_1(y)(1-u_2(x,y))=g_1(x)
  \end{equation}
  so, by\-~(\ref{g11}),
  \begin{equation}
    \int dy\ g_1(y)u_2(x,y)=0
    \label{intu0}
  \end{equation}
  and
  \begin{equation}
    g_1(x)(1+O(V^{-2}))=f_2(x)\int\frac{dy}V\ f_2(y)
    .
  \end{equation}
  Taking $\int\frac{dx}V\cdot$ on both sides, we find that
  \begin{equation}
    f_2(x)=g_1(x)(1+O(V^{-2}))
    .
  \end{equation}
\qed
\bigskip

{\bf Remark}:
Note that this proof can easily be generalized to show that $f_2=g_1(1+O(V^{-n}))$ for any $n$.

\subsubsection{Factorization of $g_3$}
\indent
We now turn to $g_3$.
\bigskip

\theo{Lemma}\label{lemma:g3}
  Assumption\-~\ref{assum:factorization} with $i=2,3$ and\-~(\ref{g11})-(\ref{g3g2}) imply that
  \begin{equation}
    g_3(x,y,z)=g_1(x)g_1(y)g_1(z)(1-u_3(x,y))(1-u_3(x,z))(1-u_3(y,z))(1+O(V^{-2}))
  \end{equation}
  with
  \begin{equation}
    u_3(x,y):=u_2(x,y)+\frac{w_3(x,y)}V
    \label{u3}
  \end{equation}
  \begin{equation}
    w_3(x,y):=(1-u_2(x,y))\int dz\ g_1(z)u_2(x,z)u_2(y,z)
    .
    \label{w3}
  \end{equation}
\endtheo
\bigskip

\indent\underline{Proof}:
  Using\-~(\ref{g3g2}) in\-~(\ref{g_factorized}),
  \begin{equation}
    g_2(x_1,x_2)=W_3(x_1,x_2)
    \int \frac{dx_3}V\ W_3(x_1,x_3)W_3(x_2,x_3)
    .
    \label{g2_factor_inproof}
  \end{equation}
  \bigskip

  \point
  We first expand to order $V^{-1}$.
  By\-~(\ref{assum_bound}),
  \begin{equation}
    \int\frac{dz}Vf_3^2(z)u_3(x,z)=O(V^{-1})
    \label{f3V1}
  \end{equation}
  so, by\-~(\ref{W_fact}),
  \begin{equation}
    g_2(x,y)=f_3^2(x)f_3^2(y)(1-u_3(x,y))
    \left(\int \frac{dz}V\ f_3^2(z)
    +O(V^{-1})\right)
    .
  \end{equation}
  By Lemma~\-\ref{lemma:g2},
  \begin{equation}
    g_1(x)g_1(y)(1-u_2(x,y))=f_3^2(x)f_3^2(y)(1-u_3(x,y))\left(\int\frac{dz}V\ f_3^2(z)+O(V^{-1})\right)
    .
  \end{equation}
  We take $\int\frac{dy}V\cdot$ on both sides of this equation.
  By\-~(\ref{intu0}) and\-~(\ref{f3V1}),
  \begin{equation}
    g_1(x)=f_3^2(x)\left(\left(\int\frac{dy}Vf_3^2(y))\right)^2+O(V^{-1})\right)
  \end{equation}
  and, integrating once more implies that $\int\frac{dy}Vf_3^2(y)=1+O(V^{-1})$.
  Therefore,
  \begin{equation}
    f_3^2(x)=g_1(x)(1+O(V^{-1}))
    \label{3fV}
  \end{equation}
  and
  \begin{equation}
    u_3(x,y)=u_2(x,y)(1+O(V^{-1}))
    .
    \label{3V}
  \end{equation}
  \bigskip

  \point
  We push the expansion to order $V^{-2}$: (\ref{g2_factor_inproof}) is
  \begin{equation}
    g_2(x,y)=f_3^2(x)f_3^2(y)(1-u_3(x,y))\int\frac{dz}{V}f_3^2(z)
    \left(
      1
      -u_3(x,z)-u_3(y,z)
      +u_3(x,z)u_3(y,z)
    \right)
    .
  \end{equation}
  By\-~(\ref{3fV})-(\ref{3V}) and Lemma\-~\ref{lemma:g2},
  \begin{equation}
    \begin{largearray}
      f_3^2(x)f_3^2(y)(1-u_3(x,y))\int\frac{dz}{V}f_3^2(z)
      =g_1(x)g_1(y)(1-u_2(x,y))
      \cdot\\[0.3cm]\hfill\cdot
      \left(1+\int\frac{dz}{V}\ (g_1(z)(u_2(x,z)+u_2(y,z)-u_2(x,z)u_2(y,z)))+O(V^{-2})\right)
      .
    \end{largearray}
  \end{equation}
  Therefore, by\-~(\ref{intu0}),
  \begin{equation}
    \begin{largearray}
      f_3^2(x)f_3^2(y)(1-u_3(x,y))\int\frac{dz}{V}f_3^2(z)=g_1(x)g_1(y)(1-u_2(x,y))
      \cdot\\\hfill\cdot
      \left(1-\int\frac{dz}{V}g_1(z)u_2(x,z)u_2(y,z)+O(V^{-2})\right)
      .
    \end{largearray}
  \end{equation}
  Now, let us apply $\int\frac{dy}V\cdot$ to both sides of the equation.
  Note that, by\-~(\ref{assum_bound}),
  \begin{equation}
    \int\frac{dy}V\ g_1(y)u_2(x,y)\int\frac{dz}Vg_1(z)u_2(x,z)u_2(y,z)=O(V^{-2})
    .
    \label{tech1}
  \end{equation}
  Furthermore, by\-~(\ref{intu0}),
  \begin{equation}
    \int \frac{dy}V\ g_1(y)u_2(x,y)=0
    ,\quad
    \int\frac{dy}V\ g_1(y)\int\frac{dz}V\ g_1(z)u_2(x,z)u_2(y,z)=0
  \end{equation}
  and by\-~(\ref{3fV}) and\-~(\ref{3V}),
  \begin{equation}
    \int\frac{dy}V\ f_3^2(y)u_3(x,y)=\int\frac{dy}V\ g_1(y)u_2(x,y)+O(V^{-2})=O(V^{-2})
    .
    \label{tech2}
  \end{equation}
  We are thus left with
  \begin{equation}
    f_3^2(x)\left(\int\frac{dy}V\ f_3^2(y)\right)^2
    =
    g_1(x)(1+O(V^{-2}))
    .
  \end{equation}
  Taking $\int\frac{dx}V\cdot$, we thus find that
  \begin{equation}
    \left(\int\frac{dx}V f_3^2(x)\right)^3=1+O(V^{-2})
  \end{equation}
  and
  \begin{equation}
    f_3^2(x)=g_1(x)(1+O(V^{-2}))
    .
  \end{equation}
  Therefore,
  \begin{equation}
    1-u_3(x,y)=(1-u_2(x,y))\left(1-\frac1V\int dz\ g_1(z)u_2(x,z)u_2(y,z)+O(V^{-2})\right)
    .
  \end{equation}
\qed

\subsubsection{Factorization of $g_4$}

\theo{Lemma}\label{lemma:g4}
  Assumption\-~\ref{assum:factorization} and\-~(\ref{g11})-(\ref{g4g2}) imply that
  \begin{equation}
    g_4(x_1,x_2,x_3,x_2)=
    g_1(x_1)g_1(x_2)g_1(x_3)g_1(x_4)
    \left(\prod_{i<j}(1-u_4(x_i,x_j))\right)
    (1+O(V^{-2}))
  \end{equation}
  with
  \begin{equation}
    u_4(x,y):=u_2(x,y)+\frac{2w_3(x,y)}V
    \label{u4}
  \end{equation}
  where $w_3$ is the same as in Lemma\-~\ref{lemma:g3}.
\endtheo
\bigskip

\indent\underline{Proof}:
  Using\-~(\ref{g4g2}) in\-~(\ref{g_factorized}),
  \begin{equation}
    g_2(x_1,g_2)=W_4(x_1,x_2)\int\frac{dx_3dx_4}{V^2}\ 
    W_4(x_1,x_3)
    W_4(x_1,x_4)
    W_4(x_2,x_3)
    W_4(x_2,x_4)
    W_4(x_3,x_4)
    .
  \end{equation}
  \bigskip

  \point
  We expand to order $V^{-1}$.
  By\-~(\ref{assum_bound}),
  \begin{equation}
    \int\frac{dz}Vf_4^3(z)u_4(x,z)=O(V^{-1})
    \label{f4V1}
  \end{equation}
  so by\-~(\ref{W_fact}),
  \begin{equation}
    g_2(x,y)=f_4^3(x)f_4^3(y)(1-u_4(x,y))\left(\int\frac{dzdt}{V^2}f_4^3(z)f_4^3(t)+O(V^{-1})\right)
    .
  \end{equation}
  By Lemma\-~\ref{lemma:g2},
  \begin{equation}
    g_1(x)g_1(y)(1-u_2(x,y))=
    f_4^3(x)f_4^3(y)(1-u_4(x,y))\left(\left(\int\frac{dz}{V}f_4^3(z)\right)^2+O(V^{-1})\right)
    .
  \end{equation}
  Applying $\int\frac{dy}V\cdot$ to both sides of the equation, using\-~(\ref{intu0}) and\-~(\ref{f4V1}),
  \begin{equation}
    g_1(x)=f_4(x)^3\left(\left(\int\frac{dy}V\ f_4^3(y)\right)^3+O(V^{-1})\right)
    .
  \end{equation}
  Integrating once more, we have $\int\frac{dy}Vf_4^3(z)=1+O(V^{-1})$ and
  \begin{equation}
    f_4^3(x)=g_1(x)(1+O(V^{-1}))
    .
    \label{4fV}
  \end{equation}
  Therefore,
  \begin{equation}
    u_4(x,y)=u_2(x,y)(1+O(V^{-1}))
    .
    \label{4V}
  \end{equation}
  \bigskip

  \point
  We push the expansion to order $V^{-2}$:
  by\-~(\ref{assum_bound}),
  \begin{equation}
    \int \frac{dzdt}{V^2}u_4(x,z)u_4(y,t)=O(V^{-2})
    ,\quad
    \int \frac{dzdt}{V^2}u_4(x,z)u_4(z,t)=O(V^{-2})
  \end{equation}
  \begin{equation}
    \int \frac{dzdt}{V^2}u_4(x,z)u_4(x,t)=O(V^{-2})
  \end{equation}
  so
  \begin{equation}
    \begin{largearray}
      g_2(x,y)=f_4^3(x)f_4^3(y)(1-u_4(x,y))
      \left(\int\frac{dzdt}{V^2}
	f_4^3(z)f_4^3(t)
	+\right.\\[0.5cm]\hfill\left.+
	\int\frac{dzdt}{V^2}
	g_1(z)g_1(t)(-2u_2(x,z)-2u_2(y,z)-u_2(z,t)+2u_2(x,z)u_2(y,z))
	+O(V^{-2})
      \right)
      .
    \end{largearray}
  \end{equation}
  By\-~(\ref{4fV}), (\ref{4V}), and Lemma\-~\ref{lemma:g2},
  \begin{equation}
    \begin{largearray}
      f_4^3(x)f_4^3(x)(1-u_4(x,y))\left(\int\frac{dz}V\ f_4^3(z)\right)^2
      =
      g_1(x)g_1(y)(1-u_2(x,y))
      \cdot\\[0.5cm]\hfill\cdot
      \left(1+
	\int\frac{dzdt}{V^2}\ g_1(z)g_1(t)(2u_2(x,z)+2u_2(y,z)+u_2(z,t)-2u_2(x,z)u_2(y,z))
	+O(V^{-2})
      \right)
      .
    \end{largearray}
  \end{equation}
  By~\-(\ref{intu0}),
  \begin{equation}
    \begin{largearray}
      f_4^3(x)f_4^3(y)(1-u_4(x,y))\left(\int\frac{dz}V\ f_4^3(z)\right)^2
      =\\[0.3cm]\hfill=
      g_1(x)g_1(y)(1-u_2(x,y))\left(1-2\int\frac{dz}{V}g_1(z)u_2(x,z)u_2(y,z)+O(V^{-2})\right)
      .
    \end{largearray}
  \end{equation}
  We apply $\int\frac{dy}V\cdot$ to both sides of the equation.
  By\-~(\ref{tech1})-(\ref{tech2}), we find
  \begin{equation}
    f_4^3(x)\left(\int\frac{dy}Vf_4^3(z)\right)^3=g_1(x)(1+O(V^{-2}))
    .
  \end{equation}
  Taking $\int\frac{dx}V\cdot$, we find that
  \begin{equation}
    f_4(x)=1+O(V^{-2})
  \end{equation}
  and
  \begin{equation}
    f_4^3(x)=g_1(x)(1+O(V^{-2}))
    .
  \end{equation}
  Therefore,
  \begin{equation}
    1-u_4(x,y)=(1-u_2(x,y))\left(1-\frac2V\int dz\ g_1(z)u_2(x,z)u_2(y,z)+O(V^{-2})\right)
    .
  \end{equation}
\qed

\subsection{Consequences of the factorization}

\point
We first rewrite\-~(\ref{eigval}) as a family of equations for $g_i$.
\bigskip

\subpoint
Integrating~\-(\ref{eigval}) with respect to $x_1,\cdots,x_N$, we find that
\begin{equation}
  E_0=
  G^{(2)}_0
  +F^{(1)}_0
  +B_0
  \label{E0}
\end{equation}
with
\begin{equation}
  G^{(2)}_0:=
  \frac{N(N-1)}{2V^2}\int dxdy\ v(x,y)g_2(x,y)
\end{equation}
\begin{equation}
  F^{(1)}_0:=
  \frac NV\int dx\ \varpi g_1(x)
\end{equation}
and $B_0$ is a boundary term:
\begin{equation}
  B_0=-\frac{N}{2V}\int dx\ \Delta g_1(x)
  .
\end{equation}
\bigskip

\subpoint
If, now, we integrate~\-(\ref{eigval}) with respect to $x_2,\cdots,x_N$, we find
\begin{equation}
  -\frac\Delta 2g_1(x)
  +\varpi g_1(x)
  +G^{(2)}_1(x)
  +G^{(3)}_1(x)
  +F^{(2)}_1(x)
  +B_1(x)
  =E_0g_1(x)
  \label{g1}
\end{equation}
with
\begin{equation}
  G^{(2)}_1(x):=\frac{N-1}V\int dy\ v(x,y)g_2(x,y)
\end{equation}
\begin{equation}
  G^{(3)}_1(x):=
  \frac{(N-1)(N-2)}{2V^2}\int dydz\ v(y,z)g_3(x,y,z)
\end{equation}
\begin{equation}
  F^{(2)}_1(x):=\frac{N-1}V\int dy\ \varpi_y g_2(x,y)
\end{equation}
in which we use the notation $\varpi_y$ to indicate that $\varpi$ applies to $y\mapsto g_2(x,y)$,
and $B_1$ is a boundary term
\begin{equation}
  B_1(x):=-\frac{N-1}{2V}\int dy\ \Delta_y g_2(x,y)
  .
\end{equation}
\bigskip

\subpoint
If we integrate with respect to $x_3,\cdots,x_N$, we find
\begin{equation}
  \begin{largearray}
    -\frac12(\Delta_x+\Delta_y)g_2(x,y)
    +v(x,y)g_2(x,y)
    +(\varpi_y+\varpi_x)g_2(x,y)
    +\\\hfill+
    G^{(3)}_2(x,y)
    +G^{(4)}_2(x,y)
    +F^{(3)}_2(x,y)
    +B_2(x,y)
    =E_0g_2(x,y)
    \label{g2}
  \end{largearray}
\end{equation}
where, here again, $\varpi_y$ indicates that $\varpi$ applies to the $y$-degree of freedom, whereas $\varpi_x$ applies to $x$,
with
\begin{equation}
  G^{(3)}_2(x,y):=
  \frac{N-2}V\int dz\ (v(x,z)+v(y,z))g_3(x,y,z)
\end{equation}
\begin{equation}
  G^{(4)}_2(x,y):=
  \frac{(N-2)(N-3)}{2V^2}\int dzdt\ v(z,t)g_4(x,y,z,t)
\end{equation}
\begin{equation}
  F^{(3)}_2(x,y):=
  \frac{N-2}V\int dz\ \varpi_z g_3(x,y,z)
\end{equation}
and $B_2$ is a boundary term
\begin{equation}
  B_2(x):=-\frac{N-2}{2V}\int dz\ \Delta_z g_3(x,y,z)
  .
\end{equation}


\point
We rewrite\-~(\ref{E0}), (\ref{g1}) and~\-(\ref{g2}) using Lemmas\-~\ref{lemma:g2}, \ref{lemma:g3} and\-~\ref{lemma:g4}.
\bigskip

\subpoint We start with\-~(\ref{E0}): by\-~(\ref{intv}) and Lemma\-~\ref{lemma:g2},
\begin{equation}
  G_0^{(2)}=
  \frac{N(N-1)}{2V^2}\int dxdy\ v(x,y)g_1(x)g_1(y)(1-u_2(x,y))+O(V^{-1})
\end{equation}
so
\begin{equation}
  E_0=
  \frac{N(N-1)}{2V^2}\int dxdy\ v(x,y)g_1(x)g_1(y)(1-u_2(x,y))
  +
  \frac NV\int dx\ \varpi g_1(x)
  +B_0
  +O(V^{-1})
  .
\end{equation}
\bigskip

\subpoint We now turn to\-~(\ref{g1}): by\-~(\ref{intv}) and Lemma\-~\ref{lemma:g2},
\begin{equation}
  G_1^{(2)}(x)=\frac{N}Vg_1(x)\left(\int dy\ v(x,y)g_1(y)(1-u_2(x,y))+O(V^{-2})\right)
\end{equation}
and by Lemma\-~\ref{lemma:g3},
\begin{equation}
  \begin{largearray}
    G_1^{(3)}(x)=
    g_1(x)\left(\frac{N^2}{2V^2}\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(x,y))(1-u_2(x,z))(1-u_3(y,z))
    -\right.\\\hfill\left.-
    \frac{3N}{2V^2}\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(y,z))
    +O(V^{-1})\right)
  \end{largearray}
\end{equation}
(we used\-~(\ref{u3}) to write $u_3=u_2+O(V^{-1})$; this works fine for $u_3(x,y)$ and $u_3(x,z)$ because the integrals over $y$ and $z$ are controlled by $v(y,z)w_3(x,y)$ and $v(y,z)w_3(x,z)$ using\-~(\ref{intv}) and\-~(\ref{assum_bound}); in the first term, it does not work for $u_3(y,z)$, as $v(y,z)w_3(y,z)$ can only control one of the integrals, and not both; the second term has an extra $V^{-1}$ that lets us replace $u_3$ by $u_2$)
and by\-~(\ref{assum_bound}) and\-~(\ref{bound_varpi}),
\begin{equation}
  F_1^{(2)}(x)=
  g_1(x)\left(\frac NV\int dy\ \varpi_y(g_1(y)(1-u_2(x,y)))
  -\frac 1V\int dy\ \varpi g_1(y)
  +O(V^{-1})
  \right)
  .
\end{equation}
The first term in $G_1^{(3)}$ is of order $V$:
\begin{equation}
  \begin{largearray}
    \frac{N^2}{2V^2}\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(x,y))(1-u_2(x,z))(1-u_3(y,z))
    =\\[0.3cm]\indent=
    \frac{N^2}{2V^2}\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(y,z))
    -
    \frac{N^2}{2V^3}\int dydz\ v(y,z)g_1(y)g_1(z)w_3(y,z)
    +\\[0.5cm]\hfill+
    \frac{N^2}{2V^2}\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(y,z))(-u_2(x,y)-u_2(x,z)+u_2(x,y)u_2(x,z))
    +O(V^{-1})
  \end{largearray}
\end{equation}
in which the only term of order $V$ is the first one, and is equal to the first term of order $V$ in $E_0$, and thus cancels out.
There is a similar cancellation between the second term of order $V$ in $F_1^{(2)}$ and $E_0$. All in all,
\begin{equation}
  \left(
    -\frac\Delta 2
    +\varpi
    +\bar G^{(2)}_1(x)
    +\bar G^{(3)}_1(x)
    +\bar F^{(2)}_1(x)
    +\bar E_0
    -B_0
  \right)
  g_1(x)
  +B_1(x)
  =g_1(x)O(V^{-1})
  \label{g1bar}
\end{equation}
with, recalling $\rho:=N/V$,
\begin{equation}
  \bar G_1^{(2)}(x):=\rho\int dy\ v(x,y)g_1(y)(1-u_2(x,y))
\end{equation}
and using\-~(\ref{w3}),
\begin{equation}
  \begin{largearray}
    \bar G_1^{(3)}(x):=
    -\frac\rho2\int \frac{dydz}V\ v(y,z)g_1(y)g_1(z)(1-u_2(y,z))\left(3+\rho \int dt\ g_1(t)u_2(y,t)u_2(z,t)\right)
    +\\[0.5cm]\hfill+
    \frac{\rho^2}2\int dydz\ v(y,z)g_1(y)g_1(z)(1-u_2(y,z))(-u_2(x,y)-u_2(x,z)+u_2(x,y)u_2(x,z))
  \end{largearray}
\end{equation}
\begin{equation}
  \bar F_1^{(2)}(x):=
  -\rho\int dy\ \varpi_y(g_1(y)u_2(x,y))
  -\int \frac{dy}V\ \varpi g_1(y)
\end{equation}
\begin{equation}
  \bar E_0:=
  \frac\rho2\int \frac{dxdy}V\ v(x,y)g_1(x)g_1(y)(1-u_2(x,y))
  .
\end{equation}
Rewriting this using\-~(\ref{avgdef})-(\ref{C}), we find\-~(\ref{compleq_g1}) with
\begin{equation}
  \Sigma_1(x):=B_1(x)-B_0g_1(x)+O(V^{-1})
  .
\end{equation}
\bigskip

\subpoint Finally, we rewrite (\ref{g2}): by\-~(\ref{intv}) and Lemma\-~\ref{lemma:g3},
\begin{equation}
  \begin{largearray}
    G_2^{(3)}(x,y)=
    \frac NVg_1(x)g_1(y)(1-u_2(x,y))
    \cdot\\\hfill\cdot
    \left(\int dz\ (v(x,z)+v(y,z))g_1(z)(1-u_2(x,z))(1-u_2(y,z))+O(V^{-1})\right)
  \end{largearray}
\end{equation}
and by Lemma\-~\ref{lemma:g4},
\begin{equation}
  \begin{largearray}
    G^{(4)}_2(x,y)=
    g_1(x)g_1(y)\left(\frac{N^2}{2V^2}(1-u_4(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_4(z,t))\Pi(x,y,z,t)
    -\right.\\\hfill\left.-
    \frac{5N}{2V^2}(1-u_2(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))
    +O(V^{-1})
    \right)
  \end{largearray}
\end{equation}
\begin{equation}
  \Pi(x,y,z,t):=
  (1-u_2(x,z))(1-u_2(x,t))(1-u_2(y,z))(1-u_2(y,t))
  \label{Pi}
\end{equation}
and by\-~(\ref{assum_bound}) and\-~(\ref{bound_varpi}),
\begin{equation}
  \begin{largearray}
    F^{(3)}_2(x,y)=
    g_1(x)g_1(y)\left(
    \frac NV(1-u_3(x,y))\int dz\ \varpi_z(g_1(z)(1-u_2(x,z))(1-u_2(y,z)))
    -\right.\\\hfill\left.-
    \frac2V(1-u_2(x,y))\int dz\ \varpi g_1(z)
    +O(V^{-1})
    \right)
    .
  \end{largearray}
\end{equation}

% Break here otherwise the glue gets overstretched
\vfill
\eject

The first term in $G_2^{(4)}$ is of order $V$: by\-~(\ref{u4}),
\begin{equation}
  \begin{largearray}
    \frac{N^2}{2V^2}(1-u_4(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_4(z,t))
    \Pi(x,y,z,y)
    =\\[0.5cm]\indent=
    \frac{N^2}{2V^2}(1-u_2(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))
    -\\[0.5cm]\indent-
    \frac{N^2}{V^3}w_3(x,y)\int dzdt\ v(z,t)g_1(z)g_1(y)(1-u_2(z,t))
    -\\[0.5cm]\indent-
    \frac{N^2}{V^3}(1-u_2(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)w_3(z,t)
    +\\[0.5cm]\indent+
    \frac{N^2}{2V^2}(1-u_2(x,y))\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))
    \left(\Pi(x,y,z,t)-1\right)
    +O(V^{-1})
  \end{largearray}
\end{equation}
in which the only term of order $V$ is the first one, and is equal to the term of order $V$ in $E_0$, and thus cancels out.
There is a similar cancellation between the term of order $V$ in $F_2^{(3)}$ and $E_0$.
All in all,
\begin{equation}
  \begin{largearray}
    \left(
      -\frac12(\Delta_x+\Delta_y)
      +v(x,y)
      +\varpi_x+\varpi_y
      +\bar G^{(3)}_2(x,y)
      +\bar G^{(4)}_2(x,y)
      +\bar F^{(3)}_2(x,y)
      +\bar E_0
      -B_0
    \right)
    \cdot\\\hfill\cdot
    g_1(x)g_1(y)(1-u_2(x,y))
    +B_2(x,y)
    =
    g_1(x)g_1(y)O(V^{-1})
  \end{largearray}
  \label{g2bar}
\end{equation}
with
\begin{equation}
  \bar G_2^{(3)}(x,y):=
  \rho\int dz\ (v(x,z)+v(y,z))g_1(z)(1-u_2(x,z))(1-u_2(y,z))
\end{equation}
and by\-~(\ref{w3}),
\begin{equation}
  \begin{array}{r@{\ }>\displaystyle l}
    \bar G_2^{(4)}(x,y)
    :=&
    -\frac\rho2\left(5+2\rho \int dr\ g_1(r)u_2(x,r)u_2(y,r)\right)\int \frac{dzdt}V\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))
    -\\[0.3cm]&-
    \rho^2\int \frac{dzdt}V\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))\int dr\ g_1(r)u_2(z,r)u_2(t,r)
    +\\&+
    \frac{\rho^2}2\int dzdt\ v(z,t)g_1(z)g_1(t)(1-u_2(z,t))
    \left(\Pi(x,y,z,t)-1\right)
  \end{array}
\end{equation}
\begin{equation}
  \begin{largearray}
    \bar F^{(3)}_2(x,y):=
    \rho\int dz\ \varpi_z(g_1(z)(-u_2(x,z)-u_2(y,z)+u_2(x,z)u_2(y,z)))
    -\\\hfill-
    \left(2+\rho\int dr\ g_1(r)u_2(x,r)u_2(y,r)\right)\int \frac{dz}V\ \varpi g_1(z)
  \end{largearray}
\end{equation}
\begin{equation}
  \bar E_0=
  \frac\rho2\int \frac{dxdy}V\ v(x,y)g_1(x)g_1(y)(1-u_2(x,y))
  .
\end{equation}
\bigskip

\subpoint
Expanding out $\Pi$, see\-~(\ref{Pi}), we find\-~(\ref{compleq_g2}) with
\begin{equation}
  \begin{array}{r@{\ }>\displaystyle l}
    \bar R_2(x,y)
    :=&
    \rho\int dz\ g_1(z)\left(
      \bar S(x,z)+\bar S(y,z)
      -2\int \frac{dt}V\ g_1(t)\bar S(t,z)
    \right)
    +\\[0.3cm]&+
    \frac{\rho^2}2\left(
      \bar S\bar\ast u_2\bar\ast u_2(x,x)+\bar S\bar\ast u_2\bar\ast u_2(y,y)
      -2\int \frac{dt}V\ g_1(t)\bar S\bar\ast u_2\bar\ast u_2(t,t)
    \right)
    +\\[0.3cm]&+
    \rho^2\int dzdt\ g_1(z)g_1(t)u_2(x,z)u_2(y,z)\left(
      \bar S(z,t)
      -\int\frac{dr}V\ g_1(r)\bar S(z,r)
    \right)
    -\\[0.3cm]&-
    \rho^2\int dt\ g_1(t)(\bar S\bar\ast u_2(x,t)+\bar S\bar\ast u_2(y,t))
    +\bar F_2^{(3)}(x,y)
    +\varpi_x+\varpi_y
  \end{array}
  \label{R1}
\end{equation}
and
\begin{equation}
  \Sigma_2(x,y):=B_2(x,y)-B_0g_1(x)g_1(y)(1-u_2(x,y))+O(V^{-1})
  .
\end{equation}
Using\-~(\ref{EA}) and\-~(\ref{C}), (\ref{R1}) becomes\-~(\ref{R}).
\bigskip

\point
Finally, (\ref{simplen}) follows from\-~(\ref{E0}) with
\begin{equation}
  \Sigma_0:=B_0+O(V^{-1})
  .
\end{equation}

\qed

\subsection{Sanity check, proof of Corollary \expandonce{\ref{cor:check}}}\label{sec:trsl_inv}
\indent
Assuming the translation invariance of the solution, $g_1(x)$ is constant.
By\-~(\ref{g11}),
\begin{equation}
  g_1(x)=1
  .
  \label{g1const}
\end{equation}
Furthermore, $\varpi\equiv 0$.
We then have
\begin{equation}
  \bar S(x,y)=S(x-y)
  ,\quad
  \bar K(x,y)=K(x-y)
  ,\quad
  \bar L(x,y)=L(x-y)
\end{equation}
(see\-~(\ref{K})-(\ref{L})).
Furthermore,
\begin{equation}
  \mathcal E(x)\equiv \mathcal E(y)\equiv\left<\mathcal E\right>=\frac\rho2\int dy\ S(y)
\end{equation}
\begin{equation}
  \bar A(x)\equiv\bar A(y)\equiv\left<\bar A\right>=\rho^2 S\ast u\ast u(0)
\end{equation}
\begin{equation}
  \bar C(x)\equiv \bar C_2(y)
  =2\rho^2\int dz\ u(z)\int dt\ S(t)
\end{equation}
which vanishes by\-~(\ref{g2g1}).
Thus,
\begin{equation}
  \bar R_2(x,y)\equiv0
  .
\end{equation}
We conclude by taking the thermodynamic limit.
\qed



\section{The momentum distribution}

\subsection{Computation of the momentum distribution, proof of Theorem \expandonce{\ref{theo:Nk}}}\label{sec:Nk_proof}
\indent
We use Theorem\-~\ref{theo:simple} with $\varpi$ as in\-~(\ref{varpiNk}).
Note that, by\-~(\ref{varpiNk}),
\begin{equation}
  \int dx\ \varpi f(x)=0
\end{equation}
which trivially satisfies\-~(\ref{bound_varpi}).
\bigskip

\point
We change variables in\-~(\ref{compleq_g2}) to
\begin{equation}
  \xi=\frac{x+y}2
  ,\quad
  \zeta=x-y
\end{equation}
and find
\begin{equation}
  \begin{largearray}
    \left(
      -\frac14\Delta_\xi-\Delta_\zeta+v(\zeta)
      -2\rho\bar K(\xi+{\textstyle\frac\zeta 2},\xi-{\textstyle\frac\zeta 2})
      +\rho^2\bar L(\xi+{\textstyle\frac\zeta 2},\xi-{\textstyle\frac\zeta 2})
      +\bar R_2(\xi+{\textstyle\frac\zeta 2},\xi-{\textstyle\frac\zeta 2})
    \right)
    \cdot\\\hfill\cdot
    g_1(\xi+{\textstyle\frac\zeta 2})g_1(\xi-{\textstyle\frac\zeta 2})
    (1-u_2(\xi+{\textstyle\frac\zeta 2},\xi-{\textstyle\frac\zeta 2}))
    =-\Sigma_2
    .
    \label{g2_xi}
  \end{largearray}
\end{equation}
In addition, by\-~(\ref{simplen}),
\begin{equation}
  e=\frac\rho2\int \frac{d\xi d\zeta}V\ g_1(\xi+{\textstyle\frac\zeta 2})g_1(\xi-{\textstyle\frac\zeta 2})v(\zeta)(1-u_2(\xi+{\textstyle\frac\zeta 2},\xi-{\textstyle\frac\zeta 2}))
  +\int\frac{dx}V\ \varpi g_1(x)
  +\Sigma_1
  .
  \label{Nken}
\end{equation}
We expand in powers of $\epsilon$:
\begin{equation}
  g_1(x)=1+\epsilon g_1^{(1)}(x)+O(\epsilon^2)
  ,\quad
  u_2(\xi+{\textstyle\frac\zeta2},\xi-{\textstyle\frac\zeta 2})=u_2^{(0)}(\zeta)+\epsilon u_2^{(1)}(\xi+{\textstyle\frac\zeta2},\xi-{\textstyle\frac\zeta 2})+O(\epsilon^2)
\end{equation}
in which we used the fact that, at $\epsilon=0$, $g_1(x)|_{\epsilon=0}=1$, see\-~(\ref{g1const}).
In particular, the terms of order $0$ in $\epsilon$ are independent of $\xi$.
Note, in addition, that, by\-~(\ref{g11}),
\begin{equation}
  \int\frac{dx}V\ g_1^{(1)}(x)=0
  .
  \label{intg11}
\end{equation}
\bigskip

\point
The trick of this proof is to take the average with respect to $\xi$ on both sides of\-~(\ref{g2_xi}).
Since we take periodic boundary conditions, the $\Delta_\xi$ term drops out.
We will only focus on the first order contribution in $\epsilon$, and, as was mentioned above, terms of order $0$ are independent of $\xi$.
Thus, the average over $\xi$ will always apply to a single term, either $g_1^{(1)}$ or $u_2^{(1)}$.
By\-~(\ref{g11}), the terms involving $g_1^{(1)}$ have zero average.
We can therefore replace $g_1^{(1)}$ by 1.
(The previous argument does not apply to the terms in which $\Delta_\zeta$ acts on $g_1$, but these terms have a vanishing average as well because of the periodic boundary conditions.)
In particular, by\-~(\ref{g2g1}) and Lemma\-~\ref{lemma:g2},
\begin{equation}
  \int\frac{d\xi}V\ (1-u_2^{(1)}(\xi+{\textstyle\frac\zeta2},\xi-{\textstyle\frac\zeta 2}))
  =1
\end{equation}
so
\begin{equation}
  \int\frac{d\xi}V\ u_2^{(1)}(\xi+{\textstyle\frac\zeta2},\xi-{\textstyle\frac\zeta 2})
  =0
\end{equation}
and thus, we can replace $u_2$ with $u_2^{(0)}$.
Thus, using the translation invariant computation detailed in Section\-~\ref{sec:trsl_inv}, we find that the average of\-~(\ref{g2_xi}) is
\begin{equation}
  (-\Delta+v(\zeta)-2\rho K(\zeta)+\rho^2 L(\zeta))(1-u_2^{(0)}(\zeta))+\epsilon F(\zeta)+O(\epsilon^2)+\Sigma_2=0
  \label{eqNk_inproof}
\end{equation}
where $K$ and $L$ are defined in\-~(\ref{K}) and\-~(\ref{L}) and $F$ comes from the contribution to $\bar R_2$ of $\varpi$, see\-~(\ref{R}):
\begin{equation}
  \begin{largearray}
    F(\zeta):=\epsilon^{-1}\int \frac{d\xi}V\ 
    \left(
      \varpi_x+\varpi_y-2\left<\varpi\right>
      +\rho\int dz\ \varpi_z(u_2^{(0)}(\xi+{\textstyle\frac\zeta2}-z)u_2^{(0)}(\xi-{\textstyle\frac\zeta 2}-z))
      -\right.\\\hfill\left.-
      \rho\int dz\ \varpi_zu_2^{(0)}(\xi+{\textstyle\frac\zeta2}-z)
      -\rho\int dz\ \varpi_zu_2^{(0)}(\xi-{\textstyle\frac\zeta 2}-z)
    \right)(1-u_2^{(0)}(\zeta))
    .
  \end{largearray}
\end{equation}
Similarly, (\ref{Nken}) is
\begin{equation}
  e=\frac\rho2\int d\zeta\ v(\zeta)(1-u_2^{(0)}(\zeta))
  +\int\frac{dx}V\ \varpi g_1(x)
  +\Sigma_1
  +O(\epsilon^2)
  .
\end{equation}
\bigskip

\point
Furthermore, by\-~(\ref{varpiNk}),
\begin{equation}
  \int dz\ \varpi_z f(z)=0
\end{equation}
for any integrable $f$, so
\begin{equation}
  F(\zeta)=\epsilon^{-1}\int \frac{d\xi}V\ 
  \left(\varpi_x+\varpi_y\right)(1-u_2^{(0)}(\zeta))
\end{equation}
and
\begin{equation}
  e=\frac\rho2\int d\zeta\ v(\zeta)(1-u_2^{(0)}(\zeta))
  +\Sigma_1
  +O(\epsilon^2)
  .
  \label{Nken_inproof}
\end{equation}
Now,
\begin{equation}
  \varpi_x f(x-y)
  =
  e^{ikx}
  \int dz\ 
  e^{-ikz}f(z-y)
\end{equation}
so
\begin{equation}
  \varpi_x f(\zeta)
  =
  \epsilon e^{ik(\xi+{\textstyle\frac\zeta2})}
  \int dz\ 
  e^{-ik(z+(\xi-{\textstyle\frac\zeta 2}))}f(z)
  =
  \epsilon e^{ik\zeta}
  \int dz\ 
  e^{-ikz}f(z)
  =\epsilon e^{ik\zeta}\hat f(-k)
  .
\end{equation}
Similarly,
\begin{equation}
  \varpi_y f(\zeta)
  =\epsilon e^{-ik\zeta}\hat f(-k)
  .
\end{equation}
Thus
\begin{equation}
  F(\zeta)=2\cos(k\zeta)(\delta(k)-\hat u_2^{(0)}(-k))
  .
  \label{F_inproof}
\end{equation}
Since $k\neq 0$, the $\delta$ function drops out.
We conclude the proof by combining\-~(\ref{eqNk_inproof}), (\ref{Nken_inproof}) and\-~(\ref{F_inproof}) and taking the thermodynamic limit.
\qed

\subsection{The simple equation and Bogolyubov theory, proof of Theorem \expandonce{\ref{theo:Nk_bog}}}\label{sec:Nk_bog}

\point
We differentiate\-~(\ref{simpleq}) with respect to $\epsilon$ and take $\epsilon=0$:
\begin{equation}
  (-\Delta+v+4e+4e\rho u\ast)\partial_\epsilon u=-4\partial_\epsilon eu+2\partial_\epsilon e\rho u\ast u+F
  .
\end{equation}
Let
\begin{equation}
  \mathfrak K_e:=(-\Delta+v+4e(1-\rho u\ast))^{-1}
\end{equation}
(this operator was introduced and studied in detail in\-~\cite{CJL21}).
We apply $\mathfrak K_e$ to both sides and take a scalar product with $-\rho v/2$ and find
\begin{equation}
  \partial_\epsilon e=\rho\partial_\epsilon e\int dx\ v(x)\mathfrak K_e(2u(x)-\rho u\ast u(x))-\frac\rho2\int dx\ v(x)\mathfrak K_eF(x)
\end{equation}
and so, using\-~(\ref{M_simpleq}),
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)=\partial_\epsilon e
  =-\frac{\frac\rho2\int dx\ v(x)\mathfrak K_eF(x)}{1-\rho\int dx\ v(x)\mathfrak K_e(2u(x)-\rho u\ast u(x))}
\end{equation}
and, by\-~(\ref{F}),
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)
  =\rho\frac{\hat u(k)\int dx\ v(x)\mathfrak K_e\cos(kx)}{1-\rho\int dx\ v(x)\mathfrak K_e(2u(x)-\rho u\ast u(x))}
  .
\end{equation}
Note that
\begin{equation}
  \int\frac{dk}{(2\pi)^3}\mathcal M^{(\mathrm{simpleq})}(k)
  =
  \frac{\rho\int dx\ v(x)\mathfrak K_e u(x)}{1-\rho\int dx\ v(x)\mathfrak K_e(2u(x)-\rho u\ast u(x))}
\end{equation}
which is the expression for the uncondensed fraction for the simple equation\-~\cite[(38)]{CHe21}.
\bigskip

\point
By\-~\cite[(5.8),(5.27)]{CJL21},
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)=\rho
  \left(\hat u(k)\int dx\ v(x)\mathfrak K_e\cos(k(x))\right)
  (1+O(\rho e^{-\frac12}))
  .
\end{equation}
Furthermore, by the resolvent identity,
\begin{equation}
  \mathfrak K_e\cos(kx)
  =
  \xi-\mathfrak K_e(v\xi)
  ,\quad
  \xi:=\mathfrak Y_e(\cos(kx))
  :=(-\Delta+4e(1-\rho u\ast))^{-1}\cos(kx)
\end{equation}
in terms of which, using the self-adjointness of $\mathfrak K_e$,
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)=\rho\hat u(k)\left(
    \int dx\ v(x)\xi(x)
    -
    \int dx\ \mathfrak K_ev(x)(v(x)\xi(x))
  \right)
  .
  \label{pde}
\end{equation}
\bigskip

\point
Now, taking the Fourier transform,
\begin{equation}
  \hat\xi(q)\equiv\int dx\ e^{ikx}\xi(x)=\frac{(2\pi)^3}2\frac{\delta(k-q)+\delta(k+q)}{q^2+4e(1-\rho\hat u(q))}
\end{equation}
and so
\begin{equation}
  \int dx\ v(x)\xi(x)
  =
  \int\frac{dq}{(2\pi)^3}\hat v(q)\hat\xi(q)
  =
  \frac{\hat v(k)}{k^2+4e(1-\rho\hat u(k))}
\end{equation}
and thus
\begin{equation}
  \rho\hat u(k)\int dx\ v(x)\xi
  =
  \rho\hat v(k)\frac{\hat u(k)}{k^2+4e(1-\rho\hat u(k))}
  .
\end{equation}
We recall\-~\cite[(4.25)]{CJL20}:
\begin{equation}
  \rho\hat u(k)=\frac{k^2}{4e}+1-\sqrt{\left(\frac{k^2}{4e}+1\right)^2-\hat S(k)}
  \label{rhou}
\end{equation}
and, by\-~\cite[(4.24)]{CJL20},
\begin{equation}
  \hat S(0)=1
  .
  \label{S1}
\end{equation}
Therefore, if we rescale
\begin{equation}
  k=2\sqrt{e}\kappa
\end{equation}
we find
\begin{equation}
  \rho\hat u(k)\int dx\ v(x)\xi
  =
  \frac{\hat v(0)}{4e}\frac{\kappa^2+1-\sqrt{(\kappa^2+1)^2-1}}{\sqrt{(\kappa^2+1)^2-1}}
  +o(e^{-1})
  .
  \label{pde1}
\end{equation}
\bigskip

\point
Now,
\begin{equation}
  \int dx\ e^{iqx}v(x)\xi(x)
  =
  \frac12\frac1{k^2+4e(1-\rho\hat u(k))}
  \int dp\ \hat v(q-p)(\delta(k-p)+\delta(k+p))
\end{equation}
so
\begin{equation}
  \int dx\ e^{iqx}v(x)\xi(x)
  =
  \frac12\frac{\hat v(q-k)+\hat v(q+k)}{k^2+4e(1-\rho\hat u(k))}
  .
\end{equation}
Therefore,
\begin{equation}
  \int dx\ \mathfrak K_ev(x)(v\xi)
  =
  \frac12\frac1{k^2+4e(1-\rho\hat u(k))}
  \int\frac{dq}{(2\pi)^3}\ 
  \widehat{\mathfrak K_e v}(q)
  (\hat v(k-q)+\hat v(k+q))
\end{equation}
which, using the $q\mapsto-q$ symmetry, is
\begin{equation}
  \int dx\ \mathfrak K_ev(x)(v\xi)
  =
  \frac1{k^2+4e(1-\rho\hat u(k))}
  \int\frac{dq}{(2\pi)^3}\ 
  \widehat{\mathfrak K_e v}(q)
  \hat v(k+q)
\end{equation}
that is,
\begin{equation}
  \rho\hat u(k)\int dx\ \mathfrak K_ev(x)(v\xi)
  =
  \frac{\rho\hat u(k)}{k^2+4e(1-\rho\hat u(k))}
  \int dx\ 
  e^{-ikx}
  \mathfrak K_e v(x)
  v(x)
\end{equation}
in which we rescale
\begin{equation}
  k=2\sqrt e\kappa
\end{equation}
so, by\-~(\ref{rhou})-(\ref{S1}),
\begin{equation}
  \rho\hat u(k)\int dx\ \mathfrak K_ev(x)(v\xi)
  =
  \frac{\kappa^2+1-\sqrt{(\kappa^2+1)^2-1}}{4e\sqrt{(\kappa^2+1)^2-1}}
  (1+o(1))\int dx\ 
  e^{-i2\sqrt e\kappa x}
  v(x)\mathfrak K_e v(x)
  .
\end{equation}
Therefore, by dominated convergence (using the argument above\-~\cite[(5.23)]{CJL21} and the fact that $\mathfrak K_e$ is positivity preserving), and by\-~\cite[(5.23)-(5.24)]{CJL21},
\begin{equation}
  \rho\hat u(k)\int dx\ \mathfrak K_ev(x)(v\xi)
  =
  \frac{\kappa^2+1-\sqrt{(\kappa^2+1)^2-1}}{4e\sqrt{(\kappa^2+1)^2-1}}
  (-4\pi a+\hat v(0))+o(e^{-1})
  .
  \label{pde2}
\end{equation}
\bigskip

\point
Inserting\-~(\ref{pde1}) and\-~(\ref{pde2}) into\-~(\ref{pde}), we find
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)
  =
  \frac{\pi a}{e}\frac{\kappa^2+1-\sqrt{(\kappa^2+1)^2-1}}{\sqrt{(\kappa^2+1)^2-1}}
  +o(e^{-1})
  .
\end{equation}
Finally, we recall\-~\cite[(1.23)]{CJL20}:
\begin{equation}
  e=2\pi\rho a(1+O(\sqrt\rho))
  \label{erho}
\end{equation}
so
\begin{equation}
  \mathcal M^{(\mathrm{simpleq})}(k)
  =
  \frac{1}{2}\frac{\kappa^2+1-\sqrt{(\kappa^2+1)^2-1}}{\sqrt{(\kappa^2+1)^2-1}}
  +o(e^{-1})
  .
  \label{final1}
\end{equation}
\bigskip

\point
Finally, by\-~(\ref{Mbog})
\begin{equation}
  \mathcal M^{(\mathrm{Bogolyubov})}(2\sqrt e\kappa)=-\frac1{2\rho}\left(1-\frac{\frac{4e}{8\pi\rho a}\kappa^2+1}{\sqrt{\frac{e^2}{4\pi^2\rho^2a^2}\kappa^4+\frac{e}{\pi\rho a} \kappa^2}}\right)
\end{equation}
so by\-~(\ref{erho}),
\begin{equation}
  \mathcal M^{(\mathrm{Bogolyubov})}(2\sqrt e\kappa)=-\frac1{2\rho}\left(1-\frac{\kappa^2+1}{\sqrt{\kappa^4+2\kappa^2}}\right)
  .
\end{equation}
This, together with\-~(\ref{final1}), implies\-~(\ref{Msimpleqbog}).
\qed

\bigskip
\bigskip

\hfil{\bf Acknowledgements}\par
The author thanks Elliott H. Lieb, Eric A. Carlen and Markus Holzmann for many valuable discussions.
The author acknowledges support from the Simons Foundation, Grant Number\-~825876.

\vfill
\eject

\begin{thebibliography}{WWW99}
\small
\IfFileExists{bibliography/bibliography.tex}{\input bibliography/bibliography.tex}{}
\end{thebibliography}


\end{document}