-
Notifications
You must be signed in to change notification settings - Fork 0
/
ch_LFT_fermions.tex
978 lines (930 loc) · 42.3 KB
/
ch_LFT_fermions.tex
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
\chapter{LFT: Lattice with Fermions}\label{ch:ferm}
We will now introduce Fermions on the lattice. Parts of this presentation
follows Chapters 5 and 10 of~\cite{gattringer_quantum_2010}.
We will start in the continuum
and introduce a naive discretization. Then we will go into detail, point
out a problem with the naive discretization, and fix it. We will omit
space-time dependence sometimes for the sake of brevity, but it should
be clear which objects have space-time dependence anyway. Primarily
we are interested in studying QCD, so $N_c=3$. This chapter uses Dirac
algebra extensively, but the algebra is different when the metric
is Euclidean. For details see \apref{ap:spec_math}.
In the continuum theory, the free fermion propagator is
\begin{equation}
S_F=\int\dd[4]{x}\bar{\psi}(\slashed{\partial}+m)\psi.
\end{equation}
Using the rules~\eqref{eq:dertodiff} and \eqref{eq:inttosum}, one naively
discretizes this action on the lattice as
\begin{equation}\label{eq:naivefermact}
S_F=a^4\sum_x\bar{\psi}(x)\left(\sum_{\mu=1}^4\gamma_\mu
\frac{\psi(x+a\hat{\mu})-\psi(x-a\hat{\mu})}{2a}
+m\psi(x)\right).
\end{equation}
The fermionic partition function is
\begin{equation}
Z_F=\int\DD{\psi}\DD{\bar{\psi}}e^{-S_F},
\end{equation}
with integration measure
\begin{equation}
\int\DD{\psi}\DD{\bar{\psi}}
=\prod_{x,f,\alpha,c}\dd{\bar{\psi}^f_{\alpha c}(x)}
\dd{\psi^f_{\alpha c}(x)},
\end{equation}
where $f$ runs over flavors, $\alpha$ runs over Dirac indices, and $c$
runs over colors.
In addition, we want the fermionic action to be gauge invariant. The
gauge transformation for fermion fields
\begin{equation}
\psi(x)\to U(x)\psi(x),~~~~~\bar{\psi}(x)\to\bar{\psi}(x)U(x)^\dagger
\end{equation}
with $U\in\SU(3)$ reminds us of the gauge transformation for scalar
fields given in Chapter~\ref{ch:preliminaries}. Working in the continuum
we introduce, just as before, a gauge field $A_\mu(x)$ so that the
fermionic action becomes
\begin{equation}
S_F=\int\dd[4]{x}\bar{\psi}\big(\gamma_\mu(\partial_\mu+A_\mu)+m\big)\psi.
\end{equation}
The gauge-transformed fermionic action is
\begin{equation}
S'_F=\int\dd[4]{x}\bar{\psi}\,U^\dagger
\big(\gamma_\mu(\partial_\mu+A'_\mu)+m\big)U\psi,
\end{equation}
so that the requirement of gauge invariance leads us to conclude
\begin{equation}
(\partial_\mu+A_\mu)\psi=U^\dagger(\partial_\mu+A'_\mu)U\psi.
\end{equation}
Solving for $A'_\mu$ we find
\begin{equation}
A'_\mu=(\partial_\mu U)U^\dagger-UA_\mu U^\dagger,
\end{equation}
just as we did before.
On the lattice, link variables transform as
\begin{equation}
U_\mu(x)\to U'_\mu(x)=U(x)U_\mu(x)U^\dagger(x+a\hat{\mu}),
\end{equation}
which ensures that closed loops of link variables are gauge invariant.
In addition the discretized fermionic action becomes gauge invariant
when it is written as
\begin{equation}\label{eq:naivefermactgauge}
S_F=a^4\sum_x\bar{\psi}(x)\left(\sum_{\mu=1}^4\gamma_\mu
\frac{U_\mu(x)\psi(x+a\hat{\mu})-U_{-\mu}(x)\psi(x-a\hat{\mu})}{2a}
+m\psi(x)\right),
\end{equation}
where
\begin{equation}
U_{-\mu}(x)=U_\mu(x-a\hat{\mu})^\dagger
\end{equation}
transforms as
\begin{equation}
U_{-\mu}(x)\to U'_{-\mu}(x)=U(x)U_{-\mu}(x)U^\dagger(x-a\hat{\mu}).
\end{equation}
Since \equatref{eq:naivefermactgauge} is the gauge-invariant action,
this will be what we investigate for fermion doubling.
\section{Grassmann
numbers}\label{sec:grassmann}\index{algebra!Grassmann}\index{Grassmann!algebra}
Fermions are objects that by definition anticommute with each other.
With this in mind, we introduce Grassmann numbers.
We consider a set of numbers $\eta_i$, $1\leq i\leq N$ obeying
\begin{equation}
\eta_i\eta_j=-\eta_j\eta_i
\end{equation}
for all $i$ and $j$. These are called {\it Grassmann
numbers}\index{Grassmann!number}.
It follows that
\begin{equation}\label{eq:gnilpotent}
\eta_i^2=0.
\end{equation}
If a power series of a function $f$ of the Grassmann numbers exists, then
\equatref{eq:gnilpotent} guarantees that the series terminates almost
immediately. In general we would write
\begin{equation}\label{eq:grasspoly}
f(\eta)=a+\sum_ia_i\eta_i+\sum_{i<j}a_{ij}\eta_i\eta_j
+...+a_{1...N}\eta_1...\eta_N
\end{equation}
with $a,\,a_i,\,a_{ij},...,\,a_{1...N}\in\mathbb{C}$.
We refer to \equatref{eq:grasspoly} as a {\it Grassmann polynomial}.
Grassmann\index{Grassmann!polynomial}
polynomials are closed under addition and multiplication,
and are thus said to form a {\it Grassmann algebra}. The $\eta_i$
are the {\it generators} of the Grassmann algebra.
To learn how differentiation ought to work, we use the simple example
$N=2$. Then
\begin{equation}
f(\eta)=a+a_1\eta_1+a_{12}\eta_1\eta_2.
\end{equation}
A definition of the derivative that follows our intuition is
\begin{equation}
\pdv{f}{\eta_1}=a_1+a_{12}\eta_2.
\end{equation}
However, the defining characteristic of Grassmann numbers is that they
anticommute. Therefore we could also have expanded $f$ as
\begin{equation}
f(\eta)=a+a_1\eta_1-a_{12}\eta_2\eta_1.
\end{equation}
In order for the derivative $f$ to make sense we must therefore require
\begin{equation}
\pdv{\eta_1}\eta_2=
-\eta_2\pdv{\eta_1}.
\end{equation}
Similarly if we take another derivative, this time with respect to $\eta_2$,
we find that partial derivatives with respect to different Grassmann variables
must also anticommute to maintain consistency. Altogether, the differentiation
rules for Grassmann variables become
\begin{equation}\begin{aligned}
\pdv{\eta_i}1&=0;\\
\pdv{\eta_i}\eta_i&=1;\\
\pdv{\eta_i}\eta_j&=
-\eta_j\pdv{\eta_i};\\
\frac{\partial^2}{\partial\eta_i\partial\eta_j}&=
-\frac{\partial^2}{\partial\eta_j\partial\eta_i}.
\end{aligned}\end{equation}
Next we move on to integration. We will construct integrals that work like
integrals over subsets $\Omega\in\mathbb{R}^N$ for which
the integrand vanishes at the boundary $\partial\Omega$.
Thus we demand that the Grassmann integral is a complex number,
\begin{equation}
\int\dd[N]{\eta}f\in\mathbb{C};
\end{equation}
that with $\lambda_1,\lambda_2\in\mathbb{C}$ it is linear,
\begin{equation}\label{eq:grasslin}
\int\dd[N]{\eta}\left(\lambda_1f_1+\lambda_2f_2\right)
=\lambda_1\int\dd[N]{\eta}f_1
+\lambda_2\int\dd[N]{\eta}f_2;
\end{equation}
and that its integrand vanishes at the boundary,
\begin{equation}\label{eq:grassBC}
\int\dd[N]{\eta}\pdv{\eta_i}f=0.
\end{equation}
If some function $f$ of $N-1$ Grassmann numbers can be written as the
derivative of another function $g$ of $N$ Grassmann numbers, it follows
from \equatref{eq:grassBC} that the integrand vanishes.
An integral of a Grassmann polynomial of $N$ variables is therefore
proportional to the coefficient $a_{1...N}$, since
$\eta_1...\eta_N$ cannot be written as a derivative of $N$ variables.
In particular if we demand a normalization
\begin{equation}\label{eq:grassnorm}
\int\dd[N]{\eta}\eta_1...\eta_N=1,
\end{equation}
we obtain the rule
\begin{equation}
\int\dd[N]{\eta}f=a_{1...N}.
\end{equation}
Finally to make Grassmann integration more analogous to the integration that
we're used to, we define
\begin{equation}
\dd[N]{\eta}\equiv\dd{\eta_N}...\dd{\eta_1}.
\end{equation}
Altogether then, the rules of Grassmann integration can be summarized by
\equatref{eq:grasslin} along with
\begin{equation}\label{eq:grassintrules}\begin{aligned}
\int\dd{\eta_i}1&=0;\\
\int\dd{\eta_i}\eta_i&=1;\\
\dd{\eta_i}\dd{\eta_j}&=-\dd{\eta_j}\dd{\eta_i}.
\end{aligned}\end{equation}
Using these properties we can define integration over subsets of Grassmann
variables. Interestingly, the measures obey the same algebraic properties as
the derivatives.
Next let us discuss how Grassmann integrals behave under linear
transformations. In general we can write such a change of variables as
\begin{equation}\label{eq:grassxform}
\eta'_i=M_{ij}\eta_j,
\end{equation}
where $M$ is a complex $N\times N$ matrix. More succinctly, one writes
$\eta'=M\,\eta$. Applying this change of variables to the normalization
\equatref{eq:grassnorm} we obtain
\begin{equation}\begin{aligned}
\int\dd[N]{\eta}\eta_1...\eta_N
&=1\\
&=\int\dd[N]{\eta'}\eta'_1...\eta'_N\\
&=\int\dd[N]{\eta'}M_{1\,i_1}...M_{N\,i_N}\,\eta_{i_1}...\eta_{i_N}\\
&=\int\dd[N]{\eta'}M_{1\,i_1}...M_{N\,i_N}\,\epsilon_{i_1...i_N}
\eta_1...\eta_N\\
&=\int\dd[N]{\eta'}\det M\,\eta_1...\eta_N.
\end{aligned}\end{equation}
It follows that under the transformation~\eqref{eq:grassxform}, the measure
transforms according to the rule
\begin{equation}
\dd[N]{\eta}=\det M\dd[N]{\eta'},
\end{equation}
which is in some sense the ``opposite" of the usual transformation rule, where
the $\det M$ would have been on the LHS.
The stuff between the parentheses in the naive action~\eqref{eq:naivefermact}
can be viewed as an operator. Keeping this in mind, we're going to show
that the fermionic partition function can be viewed as a determinant.
The setup is as follows: We have a Grassmann algebra with $2N$ generators
$\bar{\eta}_i$ and $\eta_i$ that all anticommute with each other and
an $N\times N$ linear transformation $M$.
\begin{theorem}{Matthews-Salam formula}{}
\index{Matthews-Salam formula}
$$
\int\prod_{i=1}^N\dd{\eta_i}\dd{\bar{\eta}_i}e^{\bar{\eta}M\eta}=\det M.
$$
\begin{proof} Let $\eta'=M\eta$. Then from \equatref{eq:grassxform}
we have
\begin{equation*}\begin{aligned}
\int \prod_{i=1}^N\dd{\eta_i}\dd{\bar{\eta}_i}e^{\bar{\eta}M\eta}
&=\det M\int \prod_{i=1}^N\dd{\eta'_i}\dd{\bar{\eta}_i}
\exp\left(\sum_{j=1}^N\bar{\eta}_j\eta'_j\right)\\
&=\det M\int \prod_{i=1}^N\dd{\eta'_i}\dd{\bar{\eta}_i}
\exp\left(\bar{\eta}_i\eta'_i\right)\\
&=\det M\int \prod_{i=1}^N\dd{\eta'_i}\dd{\bar{\eta}_i}
(1+\bar{\eta}_i\eta'_i)\\
&=\det M.
\end{aligned}\end{equation*}
The second line follows since pairs $\bar{\eta}_i\eta_i$ commute with
each other, the third line is a power series expansion of the second
line, and the last line follows from \equatref{eq:grassintrules}.
\end{proof}
\end{theorem}
Let us take a moment to see how the Matthews-Salam formula is employed
in lattice field theory. In our case, our Grassman variables are
fermion fields. The integration measure is a product over all
space-time sites, so that the index $i$ labels the site.
The matrix that occurs for us is the
{\it Dirac operator}\index{Dirac!operator}
or {\it Dirac matrix}\index{Dirac!matrix} $D(x,y)$, appearing
in the action like
\begin{equation}\label{eq:diracPositionBasis}
\bar{\psi}(x)\left(\slashed{D}+m\right)\psi(y)
\equiv\bar{\psi}(x)D(x,y)\psi(y).
\end{equation}
Remember that $\slashed{D}$ contains a covariant derivative, which
is responsible for parallel transport between $x$ and its neighbors,
which is why we write $D(x,y)$. When we write $D(x,y)$, we are
now thinking about $\slashed{D}+m$ as a matrix in the basis
of spacetime points. The form of $D(x,y)$ will depend on how we
discretize\footnote{For example the Wilson action, which we will discuss
in \secref{sec:LFTdoubling}, has Kronecker $\delta$-functions that
enforce each site only interacts with its immediate neighbors.
In that sense the Wilson Dirac operator is ``tridiagonal".
The HISQ action on the other hand, which we will discuss in
\secref{sec:HISQ}, includes interactions between sites and
their next-to-nearest neighbors. This is analogous to having a
band diagonal matrix.}. Hence Matthews-Salam will let us trade
away the fermion fields for the determinant of the Dirac operator.
Due to the incredible size of $D(x,y)$, computing its determinant
is prohibitively computationally expensive. We will discuss the most
commonly used trick to avoid this in \secref{sec:pseudofermions}.
If we replace $M$ with the Dirac operator, we see that the fermionic
partition function is just the determinant of the Dirac operator.
By the way, the ordering of the differentials in the measure should
be $\dd{\eta}\dd{\bar{\eta}}$. This is because the $\bar{\eta}$ variables
in the integrand will always come on the left. By placing the
$\dd{\bar{\eta}}$ differentials on the right, it will always hit the
$\bar{\eta}$ variables first.
A generalization of this gives us the generating functional
for fermions. Here we consider $4N$ Grassmann numbers $\bar{\eta}_i$,
$\eta_i$, $\bar{\theta}_i$, and $\theta_i$. The $\bar{\theta}_i$
and the $\theta_i$ are the source terms. We define
\begin{equation}\label{eq:deffermgf}
W[\theta,\bar{\theta}]\equiv\int \prod_{i=1}^N\dd{\bar{\eta}_i}\dd{\eta_i}
e^{\bar{\eta}M\eta+\bar{\theta}\eta+\bar{\eta}\theta}
\end{equation}
\begin{theorem}{}{fermgf}
$$
W[\theta,\bar{\theta}]=e^{-\bar{\theta}M^{-1}\theta}\det M.
$$
\begin{proof} Using the definition of the generating functional and
completing the square, we have
$$
\int \prod_{i=1}^N\dd{\bar{\eta}_i}\dd{\eta_i}
e^{\bar{\eta}M\eta+\bar{\theta}\eta+\bar{\eta}\theta}
=\int \prod_{i=1}^N\dd{\bar{\eta}_i}\dd{\eta_i}
e^{(\bar{\eta}+\bar{\theta}M^{-1})\,M\,
(\eta+M^{-1}\theta)-\bar{\theta}M^{-1}\theta}.
$$
Now we define new variables $\bar{\eta}'=\bar{\eta}+\bar{\theta}M^{-1}$
and $\eta'=\eta+M^{-1}\theta$. From \equatref{eq:grassintrules} it follows
that
$$
\int \prod_{i=1}^N\dd{\bar{\eta}_i}\dd{\eta_i}=
\int \prod_{i=1}^N\dd{\bar{\eta}'_i}\dd{\eta'_i}.
$$
Therefore by the Matthews-Salam formula,
\begin{equation*}\begin{aligned}
\int \prod_{i=1}^N\dd{\bar{\eta}_i}\dd{\eta_i}
e^{(\bar{\eta}+\bar{\theta}M^{-1})\,M\,
(\eta+M^{-1}\theta)-\bar{\theta}M^{-1}\theta}
&=\int \prod_{i=1}^N\dd{\bar{\eta}'_i}\dd{\eta'_i}
e^{\bar{\eta}'M\eta'-\bar{\theta}M^{-1}\theta}\\
&=e^{-\bar{\theta}M^{-1}\theta}\det M.
\end{aligned}\end{equation*}
\end{proof}
\end{theorem}
With the knowledge we now have, we can derive Wick's theorem, which lets us
calculate fermionic expectation values. First we define
\begin{equation}\label{eq:deffermev}
\ev{\eta_{i_1}\bar{\eta}_{j_1}...\eta_{i_n}\bar{\eta}_{j_n}}_F
\equiv\frac{1}{Z_F}\int\prod_{k=1}^N\dd{\eta_k}\dd{\bar{\eta}_k}
\bar{\eta}_{j_1}...\eta_{i_n}\bar{\eta}_{j_n}
e^{-\bar{\eta}M\eta}.
\end{equation}
\begin{theorem}{Wick's theorem}{}
\index{Wick's theorem}
$$
\ev{\eta_{i_1}\bar{\eta}_{j_1}...\eta_{i_n}\bar{\eta}_{j_n}}_F
=(-1)^n\sum_{P_{1,...,n}}\sign P\,
M^{-1}_{i_1,j_{P_1}}\,...\,M^{-1}_{i_n,j_{P_n}}.
$$
\begin{proof} From the definitions~\eqref{eq:deffermgf} and
\eqref{eq:deffermev} we have
$$
\ev{\eta_{i_1}\bar{\eta}_{j_1}...\eta_{i_n}\bar{\eta}_{j_n}}_F
=\frac{1}{Z_F}
\pdv{\theta_{j_1}}
\pdv{\bar{\theta}_{i_1}}\,...\,
\pdv{\theta_{j_n}}
\pdv{\bar{\theta}_{i_n}}W[\theta,\bar{\theta}]
\Big|_{\theta=\bar{\theta}=0}.
$$
Let us now apply Theorem~\ref{thm:fermgf} to the RHS and carry
out the derivatives. The $\det M$ will cancel with $Z_F$ because
of the Matthews-Salam formula.
\end{proof}
\end{theorem}
\section{Fermion doubling}\label{sec:LFTdoubling}
Let us now discuss one of the important problems with the naive fermionic
action. First we introduce some results about Fourier transformations on the
lattice. Define
\begin{equation}
V\equiv N_1N_2N_3N_4
\end{equation}
with $N_\mu$ even.\index{BCs!toroidal}
We generalize to {\it toroidal BCs}, i.e.
\begin{equation}
f(x+aN_\mu\hat{\mu})=e^{2\pi i\theta_\mu}f(x).
\end{equation}
Periodic BCs then have $\theta_\mu=0$ and {\it anti-periodic BCs} have
\index{BCs!anti-periodic}
$\theta_\mu=1/2$. The momentum space becomes
\begin{equation}
p_\mu=\frac{2\pi}{aN_\mu}(k_\mu+\theta_\mu),~~~~~~
-\frac{N_\mu}{2}<k_\mu\leq\frac{N_\mu}{2},
\end{equation}
which reduces to the first Brillouin zone when periodic BCs are employed. By
including the boundary phases in the momentum definition, plane waves
\begin{equation}
\exp(ip_\mu x_\mu)
\end{equation}
will also obey the BCs. Now we derive a basic formula for Fourier
transformations on the lattice. Let $N$ be even and $\ell$ be an integer
$0\leq\ell\leq N-1$.
\begin{proposition}{}{}
$$
\frac{1}{N}\sum_{j=-N/2+1}^{N/2}\exp(\frac{2\pi i\ell}{N})^j=\delta_{\ell 0}.
$$
\begin{proof} Since $N$ is even there are $N$ terms in the above sum,
so we find the LHS to be 1. For $\ell\neq 0$ let
$m\equiv j+N/2+1$ and define
$$
q\equiv\exp(\frac{2\pi i\ell}{N}).
$$
Then
\begin{equation*}\begin{aligned}
\frac{1}{N}\sum_{j=-N/2+1}^{N/2}\exp(\frac{2\pi i\ell}{N})
\propto\sum_{m=0}^{N-1}q^m
=\frac{1-q^N}{1-q}=0
\end{aligned}\end{equation*}
since $q^N=1$.
\end{proof}
\end{proposition}
Applying this formula in each space-time direction, we arrive at
\begin{equation}
\frac{1}{V}\sum_p\exp\big(ip_\mu(x-x')_\mu\big)
=\delta(x-x')
=\delta_{n_1n'_1}\delta_{n_2n'_2}\delta_{n_3n'_3}\delta_{n_4n'_4},
\end{equation}
where $x_\mu=an_\mu$, and
\begin{equation}
\frac{1}{V}\sum_x\exp\big(i(p-p')_\mu x_\mu\big)
=\delta(p-p')
=\delta_{k_1k'_1}\delta_{k_2k'_2}\delta_{k_3k'_3}\delta_{k_4k'_4}.
\end{equation}
We define the Fourier transform as
\begin{equation}\label{eq:latft}
\tilde{f}(p)=\frac{1}{\sqrt{V}}\sum_xf(x)\exp(-ip_\mu x_\mu).
\end{equation}
The inverse transform
\begin{equation}\label{eq:latift}
f(x)=\frac{1}{\sqrt{V}}\sum_p\tilde{f}(p)\exp(ip_\mu x_\mu)
\end{equation}
can be verified by plugging the Fourier transform into the LHS.
Using these results about Fourier transformations on the lattice, we
can begin to investigate fermion doubling. We shall do this for
a single flavor for notational convenience; the result clearly
generalizes to a summation over flavors.
The gauge-invariant, naive
fermion action~\eqref{eq:naivefermactgauge} is bilinear in $\psi$
and $\bar{\psi}$, so it can be written as
\begin{equation}
S_F=a^4\sum_{x,y}\sum_{\alpha,\beta,c_1,c_2}\bar{\psi}(x)_{\alpha c_1}
\,D(x,y)_{\alpha \beta c_1 c_2}\,\psi(y)_{\beta c_2},
\end{equation}
where $\alpha$ and $\beta$ are Dirac indices, $c_1$ and $c_2$ are
color indices, and the Dirac operator on the lattice is
\begin{equation}
D(x,y)_{\alpha\beta c_1c_2}\equiv\sum_{\mu=1}^4(\gamma_\mu)_{\alpha\beta}
\frac{U_\mu(x)_{c_1c_2}\delta_{x+a\hat{\mu},y}
-U_{-\mu}(x)_{c_1c_2}\delta_{x-a\hat{\mu},y}}{2a}
+m\delta_{\alpha\beta}\delta_{c_1c_2}\delta_{xy}.
\end{equation}
In this form, we can apply the Matthews-Salam formula or Wick's
theorem by identifying $M=a^4D$.
For simplicity, we set $U_\mu(x)=\id$
for all links, just so we can see clearly how the doubling arises.
Since we have free fermions, we may as well suppress color indices,
and we will also use vector/matrix notation in Dirac space, so that
Dirac indices are also suppressed. We find
\begin{equation}\begin{aligned}
\tilde{D}(p,q)&=\frac{1}{V}\sum_{x,y}e^{-ipx}D(x,y)e^{-iqy}\\
&=\frac{1}{V}\sum_{x,y}e^{-ipx}\left(
\sum_\mu\gamma_\mu\frac{\delta_{x+a\hat{\mu},y}
-\delta_{x-a\hat{\mu},y}}{2a}+m\delta_{x,y}\id
\right)e^{-iqy}\\
&=\frac{1}{V}\sum_xe^{-i(p+q)x}\left(
\sum_\mu\gamma_\mu\frac{e^{iq_\mu a}-e^{-iq_\mu a}}{2a}+m\id\right)\\
&=\delta(p+q)\left(\frac{i}{a}\sum_\mu
\gamma_\mu\sin(q_\mu a)+m\id\right)\\
&\equiv\delta(p+q)\tilde{D}(q).
\end{aligned}\end{equation}
In Ref.~\cite{gattringer_quantum_2010} they take the transform
of the $y$ part to have opposite sign, which they say makes the similarity
transformation unitary. For the purpose of seeing fermion doubling
the sign does not make a difference; it only changes the delta
function from $\delta(p-q)$ to $\delta(p+q)$. Therefore I decided to stick
with the convention~\eqref{eq:latft}. If we want to use Wick's theorem,
we will need to calculate the inverse of the Dirac operator. We find
\begin{equation}\label{eq:latprop}
\tilde{D}(p)^{-1}=\frac{m\id-ia^{-1}\sum_\mu\gamma_\mu\sin(p_\mu a)}
{m^2+a^{-2}\sum_\mu\sin(p_\mu a)^2},
\end{equation}
which can be verified by multiplying both sides by $\tilde{D}(p)$. This
is the propagator for free fermions in momentum space. At $m=0$ we find
as $a\to0$
\begin{equation}
\tilde{D}(p)^{-1}\to\frac{-i\sum_\mu\gamma_\mu p_\mu}{p^2}.
\end{equation}
This propagator has a pole at $p=0$. Poles in propagators correspond
to physical particles, so in the continuum theory, the propagator
corresponds to a single particle satisfying the Dirac equation. On
the lattice if the fermions are massless, \equatref{eq:latprop} has
a pole anywhere $p_\mu a=\pi$. In the first Brillouin zone, this
happens 15 places besides $p_\mu=0$. Hence on the lattice at finite
spacing, the propagator has 15 unphysical poles that nevertheless
correspond to some fermions. This is called {\it fermion doubling},
\index{fermion!doubling}
and we call the 15 unwanted particles the {\it doublers}.
This proliferation of extra particles can cause problems in the
interacting theory. The additional states can be pair produced
through interactions of the fermion field~\cite{montvay_quantum_1994}.
For instance even if all particles on the external lines of a diagram
are the real particles, the doublers can appear in virtual loops.
Therefore one typically wants to remove the doublers.
One way to remove the doublers is to introduce an extra term that
cancels the doublers on the lattice, and still reduces to the
correct continuum value. With this formulation, the lattice
momentum space Dirac operator is
\begin{equation}
\tilde{D}(p)=m\id+\frac{i}{a}\sum_\mu\gamma_\mu\sin(p_\mu a)
+\frac{1}{a}\sum_\mu\id\big(1-\cos(p_\mu a)\big).
\end{equation}
This second term is called the {\it Wilson!term}. This formulation
\index{fermion!Wilson}\index{Wilson!fermions}
of fermions is called {\it Wilson fermions}. The complete Dirac
operator using this formulation becomes
\begin{equation}
D(x,y)^f_{\alpha\beta c_1c_2}
=\left(m^f+\frac{4}{a}\right)\delta_{\alpha\beta}
\delta_{c_1c_2}
\delta_{xy}
-\frac{1}{2a}\sum_{\mu=\pm 1}^{\pm 4}(\id-\gamma_\mu)_{\alpha\beta}
U_\mu(x)_{c_1 c_2}\delta_{x+a\hat{\mu},y},
\end{equation}
where
\begin{equation}
\gamma_{-\mu}\equiv-\gamma_{\mu}.
\end{equation}
\section{Staggered
fermions}\index{fermion!staggered}\index{staggered!fermions}\label{sec:stagg}
In this section we work in a free theory to simplify the notation.
In the interacting theory, each spinor will have a link attached to it
originating at the same point, and the Dirac indices that we will
manipulate commute with these links, so the derivation will be the same.
This follows parts of Section 10.1 of Ref.~\cite{gattringer_quantum_2010}
and Section 4.4 in Ref.~\cite{rothe_lattice_2005}.
If you would like to take a deeper look at staggered fermions,
Ref.~\cite{Golterman:2024xos} gives a comprehensive review.
In \secref{sec:LFTdoubling} we saw that the doubling problem could
be alleviated by adding a Wilson term of the form
\begin{equation}
\frac{1}{a}\sum_\mu\id\big(1-\cos(p_\mu a)\big).
\end{equation}
Writing this as exponentials, and doing the Fourier back transformation,
we can express this as
\begin{equation}
a\sum_{\mu=1}^4\frac{1}{2a^2}
\left(2\delta_{x,y}-\delta_{x,y-a\hat{\mu}}-\delta_{x,y+a\hat{\mu}}\right).
\end{equation}
In this form, one clearly sees that the Wilson term vanishes in the
naive continuum limit like $a$. Moreover it has the form of a
discretized $\partial_\mu\partial_\mu$. A drawback of this term,
however, is that it breaks $\SU_A(N_f)\times\U_A(1)$ explicitly
because the Wilson term commutes with $\gamma_5$.
For some calculations one may desire not to break this symmetry entirely,
in particular when one is investigating phenomena closely related to
chiral symmetry. This motivates other possible fermion discretizations
that remove doublers. {\it Staggered fermions} will remove some of the
unphysical doublers, while keeping a remnant of this axial symmetry intact.
This is accomplished by redistributing the quark degrees of freedom over
the lattice in such a way that the spacing between each degree of freedom
doubles, thereby doubling the effective lattice spacing, and hence
reducing the size of the Brillouin zone. New quarks called {\it tastes}\index{taste}
will be constructed by linear combinations of these degrees of freedom.
We take as the starting point the naive fermion action
\begin{equation}
S_F
=a^4\sum_n\bar{\psi}(n)\left(\sum_{\mu=1}^4\gamma_\mu
\frac{\psi(n+\hat{\mu})-\psi(n-\hat{\mu})}{2a}
+m\psi(n)\right).
\end{equation}
In this equation and in what follows I will write the space-time coordinate
as $n$ to make the connection between the space-time coordinate and the
following more clear. The {\it staggered
transformation}\index{staggered!transformation} is given by
\begin{equation}\begin{aligned}
\psi(n)&=\gamma_1^{n_1}\gamma_2^{n_2}\gamma_3^{n_3}\gamma_4^{n_4}\psi(n)'\\
\bar{\psi}(n)
&=\bar{\psi}(n)'\gamma_4^{n_4}\gamma_3^{n_3}\gamma_2^{n_2}\gamma_1^{n_1}
\end{aligned}\end{equation}
The fact that these transformed fields come with power of gamma matrices,
along with the fact that $\gamma_\mu^2=\id$, will
diagonalize the naive action in Dirac space. In particular one finds
relationships such as
\begin{equation}
\bar{\psi}(n)\gamma_4\psi(n+\hat{4})
=(-1)^{n_1+n_2+n_3}\bar{\psi}(n)'\id\psi(n+\hat{4}),
\end{equation}
which means that in the kinetic part, the gamma matrices are annihilated,
and what remains are scalar {\it staggered phases}\index{staggered!phase}
\begin{equation}\label{eq:staggeredPhase}
\eta_1(n)=1, ~~~
\eta_2(n)=(-1)^{n_1}, ~~~
...\,, ~~~
\eta_4(n)=(-1)^{n_1+n_2+n_3}.
\end{equation}
Substituting the staggered fields
into the naive action, we thus find
\begin{equation}\label{eq:SFdiag}
S_F
=a^4\sum_n\bar{\psi}(n)'\id\left(\sum_{\mu=1}^4\eta_\mu(n)
\frac{\psi(n+\hat{\mu})'-\psi(n-\hat{\mu})'}{2a}
+m\psi(n)'\right),
\end{equation}
which is now diagonal in Dirac space because the gammas are gone.
The Dirac operator of \equatref{eq:SFdiag} represents four copies of
the same equations along its diagonal, so we can throw away three copies
without losing any information. We will call this last, kept copy $\chi$;
another way of looking at this is we keep only one Dirac component.
With this notation, we find a particularly simple form for the staggered
action:
\begin{equation}\label{eq:stagg1}
S_F^{\rm stag}
=a^4\sum_n\bar{\chi}(n)\left(\sum_{\mu=1}^4\eta_\mu(n)
\frac{\chi(n+\hat{\mu})-\chi(n-\hat{\mu})}{2a}
+m\chi(n)\right)
\end{equation}
Equation \eqref{eq:stagg1} is convenient for simulations because it contains no
Dirac spinors explicitly, but we would like a form that tells us a little
more about the physics. We will now begin to rewrite this equation in such
a way that we see the increase of the effective lattice spacing. The
strategy will be to divide the lattice into disjoint unit hypercubes.
The $\chi$ live on the corners of the hypercubes. New fields $q$ will be
linear combination of the $\chi$, and their sites will be indexed by
hypercube number.
To this end, we introduce new vectors $h$ (hypercube) and $s$ (corner) that
are related to the original site vectors by
\begin{equation}\label{eq:hypercubeIndex}
n_\mu=2h_\mu+s_\mu~~~\text{with}~~~h_\mu\in\{0, 1, ..., N_\mu/2\}
~~~s_\mu\in\{0,1\}.
\end{equation}
Moving in the direction $\mu$, $h_\mu$ counts the hypercube number,
while $s$ is a displacement vector starting at the hypercube's origin.
In this indexing scheme
\begin{equation}
\eta_\mu(n)=\eta_\mu(2h+s)=\eta_\mu(s)
\end{equation}
because contributions to powers of gamma coming from the hypercubes are
always even, according to \equatref{eq:hypercubeIndex}.
Next we introduce $\Gamma$, which will serve as the weights in the linear
combination. We define
\begin{equation}
\Gamma^s=\gamma_1^{s_1}\gamma_2^{s_2}\gamma_3^{s_3}\gamma_4^{s_4}.
\end{equation}
These obey, respectively, orthogonality and completeness
relations\footnote{The orthogonality relation can be proven using
properties of the Euclidean $\gamma$ matrix, in particular that
$\gamma_\mu=\gamma_\mu^\dagger=\gamma_\mu^{-1}$.}
\begin{equation}
\frac{1}{4}\tr\left[\Gamma^{s\dagger}\Gamma^{s'}\right]=\delta_{ss'}
~~~~\text{and}~~~~
\frac{1}{4}\sum_s\Gamma_{ba}^{s*}\Gamma_{b'a'}^s=\delta_{aa'}\delta_{bb'}
\end{equation}
Our new quark fields are then (see \figref{fig:staggHypercubes})
\begin{equation}\label{eq:staggqdef}
q(h)_{ab}\equiv\frac{1}{8}\sum_s\Gamma^s_{ab}\chi(2h+s)
~~~~\text{and}~~~~
\bar{q}(h)_{ab}\equiv\frac{1}{8}\sum_s\bar{\chi}(2h+s)\Gamma^{s*}_{ab}.
\end{equation}
In order to cast the staggered action in terms of $q(h)$, we need to
invert \equatref{eq:staggqdef}. This can be done using the
orthogonality relation. One obtains
\begin{equation}
\chi(2h+s)=2\tr\left[\Gamma^{s\dagger}q(h)\right]
~~~~\text{and}~~~~
\bar{\chi}(2h+s)=2\tr\Big[\bar{q}(h)\Gamma^s\Big].
\end{equation}
\begin{figure}[t]
\centering
\includegraphics{figs/staggHypercubes.pdf}
\caption{Example subdivision of one direction into disjoint hypercubes.
Each hypercube is labeled by $h$, and the field $q(h)$
associated to the hypercube is a linear combination of the
$\chi$ fields on the corners, here represented by
different colors. The effective lattice spacing between
a $\chi$ of one corner of the blue hypercube and the corresponding
$\chi$ in the same corner of the red hypercube is $a'=2a$.
}
\label{fig:staggHypercubes}
\end{figure}
We are now ready to carry out the substitution. The mass term follows
straightforwardly from the completeness relation. The kinetic term is
difficult, however, because the site offsets mix hypercubes. Therefore
it is crucially important to keep track of which hypercube $\chi$
belongs to. Keeping track of the hypercube will make manifest an $s$
dependence that makes applying the completeness relation not possible,
at least superficially. Gattringer and Lang~\cite{gattringer_quantum_2010}
suggest the following trick: we can shift the site that the first
hypercube starts on by one, which gives equivalent results as the
original hypercube labeling convention, and then average over these
two conventions. I was not able to figure this out.
Rothe~\cite{rothe_lattice_2005} has another strategy, which I was also
not able to follow through. Nevertheless, one should find
\begin{equation}
S_F^{\rm stag}=a'^4\sum_h\Big(m\tr\left[\bar{q}q\right]
+\tr\left[\bar{q}\gamma_\mu\partial_\mu q\right]
-\frac{a'}{2}\tr\left[\bar{q}\gamma_5\Box_\mu q\gamma_\mu\gamma_5\right]\Big)
\end{equation}
where we have an implied summation over $\mu$, we have written
$q=q(h)$, introduced the effective spacing $a'\equiv2a$, and
the discretized derivatives are
\begin{equation}
\partial_\mu f(h)\equiv\frac{f(h+\hat{\mu})-f(h-\hat{\mu})}{2a'}
\end{equation}
and
\begin{equation}
\Box_\mu f(h)\equiv\frac{f(h+\hat{\mu})-2f(h)+f(h-\hat{\mu})}{a'^2}.
\end{equation}
Our final step in the exploration of the staggered action will be to
identify the unphysical fermions. These {\it tastes} are hidden in
\index{taste}
the $q$. This should not be surprising because the $q_{ab}$ have 16
components, but Dirac spinors should have 4. This
tells us each $q$ corresponds to 4 spinors, and we will identify from
the $a$ and $b$ a taste index and Dirac index.
By comparing e.g. the $\tr[\bar{q}q]=\bar{q}_{ab}q_{ba}$ term with
what we expect from the physical, continuum theory, it makes sense
to identify the Dirac index with $b$. Thus our taste spinors are
\begin{equation}
\psi^{t}(h)_\alpha\equiv q(h)_{\alpha t}~~~~\text{and}~~~~
\bar{\psi}^{t}(h)_\alpha\equiv \bar{q}(h)_{t\alpha}.
\end{equation}
The staggered fermion action then becomes
\begin{equation}\label{eq:staggPhys}
S_F^{\rm stag}=a'^4\sum_h\left(
m\bar{\psi}^{t}\psi^{t}
+\bar{\psi}^{t}\gamma_\mu\partial_\mu\psi^{t}
-\frac{a'}{2}\bar{\psi}^{t}\gamma_5(\tau_5\tau_\mu)_{tt'}\Box_\mu\psi^{t'}
\right),
\end{equation}
where we have introduced new matrices
\begin{equation}
\tau_\mu\equiv\gamma_\mu^T.
\end{equation}
With the staggered action in the form of \equatref{eq:staggPhys} we can
begin to discuss some physics. The last term is called the
\index{taste!breaking}
{\it taste-breaking} term. It is similar to the Wilson term, in the
sense that it also represents a second derivative. However unlike
the Wilson term, it allows for interactions between fermions of
\index{taste!mixing}
different taste, i.e. it allows for {\it taste mixing}. If not for
the taste-breaking term, fermions of different taste would be mass-degenerate,
which one sees clearly from the first term. In the naive continuum limit,
this taste breaking terms vanishes like $a$.
The Wilson term broke axial symmetry completely, but in the taste-breaking
term, the remnant $\U_A(1)\times\U_A(1)$ remains. In particular this
term is invariant under transformations
\begin{equation}
\psi'=e^{i\omega}\psi, ~~~~~~
\bar{\psi}'=\bar{\psi}e^{-i\omega}
\end{equation}
and
\begin{equation}
\psi'=e^{i\omega\gamma_5\otimes\tau_5}\psi, ~~~~~~
\bar{\psi}'=\bar{\psi}e^{i\omega\gamma_5\otimes\tau_5}.
\end{equation}
This latter symmetry follows from the fact that $\gamma_5$ commutes through
the taste-breaking term, while $\tau_5$ will pick up a minus sign.
One can identify this symmetry with a subgroup of the axial taste
symmetry group $\SU_A(N_t)$, where $N_t$ is the number of tastes.
At finite lattice spacing, taste mixing lifts the taste mass degeneracy.
One way to get some feeling for taste-breaking effects, then, is to look
at the mass spectrum of staggered fermions. It has been found that these
taste breaking effects can be reduced by improved gauge actions or smearing
\cite{durr_staggered_2004,follana_index_2004}, and in particular smearing
seems to drive masses to 4-fold degeneracy. An intuition for why smearing
might help is as follows: In the interacting theory, each $\chi$ is attached
to links according to its site, and since tastes are linear combinations
of these, it follows that different tastes touch different links. So
the more ``distance" in $\SU(N)$ space between the links, i.e. the more
the links fluctuate, the greater the taste-breaking effects will be.
Since smearing algorithms tend to drive links to more typical values
given their neighbors, they reduce these fluctuations, and hence
the taste-breaking.
We would also like to suppress the effects of unphysical tastes. Absent
taste-breaking, the Dirac operator corresponding to the staggered action
would be block diagonal in taste space, i.e. we would have for
one physical flavor
\begin{equation}
D^{\rm stag}=\left(\begin{array}{cccc}
D & & & \\
& D & & \\
& & D& \\
& & & D\\
\end{array}\right).
\end{equation}
One commonly used strategy to remove the effects of taste-breaking is
\index{staggered!rooting}
therefore {\it rooting}. The idea is that $\det D^{\rm stag}$ is
the contribution to the probability distribution from four mass-degenerate
flavors. To isolate one of the flavors, it is sufficient to use
$\det D=(\det D^{\rm stag})^{1/4}$. For two degenerate light flavors
$m_l$ and one heavier flavor $m_s$, i.e. for $N_f=2+1$ fermions,
one then samples with probability distribution
\begin{equation}\label{eq:HISQdist}
\dd{\rm P}
=\dd Ue^{-S_G}(\det D_l^{\rm stag})^{1/2}(\det D_s^{\rm stag})^{1/4}.
\end{equation}
In practice, taste-breaking is present at each lattice spacing, and
therefore it is not clear whether there is some leftover effect of rooting
in the continuum limit. In spite of this danger, people who employ staggered
actions often tend to use rooting anyway. One way to make this step more
justified is to reduce taste-breaking effects. Improved
actions\footnote{We will discuss improved actions in more detail in
\chref{ch:improve}.} such
as the HISQ action, to be discussed in the next section, have greatly
reduced taste-breaking and, reassuringly, results from HISQ actions
seem to agree with experiment.
%\section{Overlap fermions}
\section{Constructing interpolators}
One of the goals of LQCD calculations is to learn something about hadrons. A
lattice operator that is supposed to represent some particular observable is
called an {\it interpolator}\index{interpolator}. Hence to learn something
about hadrons, we need to construct their corresponding interpolators. The
guiding principles to constructing them are:
\begin{enumerate}
\item They are local, i.e. they should create/annihilate the particle of
interest at a particular space-time point.
\item They should be constructed from the quark fields that make the particle
of interest.
\item They need to have quantum numbers matching the particle of interest,
e.g. they need to have the correct isospin, parity, electric charge,
etc.
\end{enumerate}
As an example, let's take the\index{interpolator!pion} pions. They are known to
have\footnote{These properties are determined experimentally. For
example the pion parity was found in the 1950s~\cite{chinowsky_reaction_1955}.}
the quantum numbers $J=0$, $P=-1$, and $I=1$. There are three
pions, $\pi^{\pm}$ and $\pi^0$ with electric charges $\pm1$ and 0,
respectively; thus according to \tabref{tab:discreteSymm},
\secref{sec:isohyper}, and the above rules, we must have\footnote{The relative
minus sign in $\pi^0$ is to get $I_z=0$.}
\begin{equation}\begin{aligned}
\pi^+(x) &= \bar{d}(x)\gamma_5u(x),\\
\pi^-(x) &= \bar{u}(x)\gamma_5d(x),\\
\pi^0(x) &= \frac{1}{\sqrt{2}}\left( \bar{u}(x)\gamma_5u(x)
-\bar{d}(x)\gamma_5d(x)\right).
\end{aligned}\end{equation}
These interpolators correspond to the pion operators in our physical Hilbert
space; they annihilate a pion at $x$. In general, meson interpolators have the
form
\begin{equation}\label{eq:mesonInterp}
X=\bar{\psi}^f\Gamma\psi^g
\end{equation}
for some quarks $f$ and $g$. It turns out that to create meson $X$ one needs
\begin{equation}
\bar{X}=\bar{\psi}^g\Gamma\psi^f.
\end{equation}
With the interpolators in hand, we can construct general meson correlators.
A correlation function\index{correlator}
that creates a meson $X$ at $y$ and annihilates it at $x$ is
\begin{equation}
G(x,y)\equiv\bra{0}X(x)\bar{X}(y)\ket{0}=\ev{X(x)\bar{X}(y)}.
\end{equation}
% the fields become D^-1
% depending on how the contractions go, you can get connected and
% disconnected pieces, GL 6.1.2
% definite momentum requires momentum projection (fourier xform),
% which according to LGT:prelim gives sum over \vec{x}.
% is the discussion from degrand-detar more general?
% section 6.2.3 wall sources
\section{Pseudofermions}\label{sec:pseudofermions}
As discussed in \secref{sec:grassmann}, the Matthews-Salam formula
lets us integrate out the fermion fields, and we are left with
the determinant of a very large matrix, the Dirac matrix, in
our integration measure. Computing this determinant directly
is not computationally feasible. Let's follow a nice presentation
from Ref.~\cite{degrand_lattice_2006} to learn how to approach
simulating dynamical fermions in spite of this difficulty.
The partition function for a single fermion species is
\begin{equation}\begin{aligned}
Z&=\int\DD U\DD\bar\psi\DD\psi e^{-S_G-\bar\psi D\psi}
&=\int\DD U \det D\,e^{-S_G}.
\end{aligned}\end{equation}
As we can see in \eqref{eq:diracPositionBasis}, we think of $D$ as a matrix in
the basis of space-time sites, i.e. there is an element for each pair of
space-time points. Even though it is a sparse matrix, its dimensionality is
still impossibly huge, and hence computing determinants in a straightforward
manner is computationally not feasible\footnote{As discussed with
staggered fermions, we will want to take
square roots and fourth roots of the Dirac operator. These rooted matrices
have even more nonzero entries, exacerbating the problem significantly.
It should be that the further one gets from the diagonal, the smaller
the entries become, and in the continuum limit, $D$
should become local\index{local}. It should be noted that
the time complexity of
computing the determinant of an $N\times N$ matrix
is usually $\order{N^3}$. This is of course algorithm-dependent,
but the lowest I am aware of is still
slower than $\order{N^2}$, which for us would correspond to
the lattice volume squared.}.
Instead of evaluating the determinant, we approximate it
by introducing
{\it pseudofermions}~\cite{fucito_proposal_1981,weingarten_physics_1981}.
A pseudofermion $\Phi\equiv\Phi_R+i\Phi_I$ is a color-triplet, complex,
scalar field. We leverage the formal identity
\begin{equation}\label{eq:pseudofermionIdentity}
\det M \sim \int\DD{\Phi_R}\DD{\Phi_I}\exp\left(-\Phi^\dagger M^{-1}\Phi\right),
\end{equation}
where the squiggle indicates there may be some irrelevant factors
missing\footnote{Remember that the only important thing when it comes to
predicting observables is that expectation values remain unchanged.
Constant factors always drop out in the ratio. In Gattringer and Lang I
see some factors $\pi^N$, which I do not always see in the literature.}.
Now, it is always crucial that our exponentials can be interpreted
as probabilities; therefore an important question to ask is whether
that exponential is real and bounded.
One way to explore the argument of the above exponential is to decompose
the scalar fields $\Phi$ in terms of eigenvectors of $M$. Assuming $M$
is an $N\times N$ matrix with only nonzero eigenvalues $\lambda_i$,
we can write
\begin{equation}
\Phi^\dagger M^{-1}\Phi=\sum_{i=1}^N
\left(\Phi^\dagger\psi\right)_i\frac{1}{\lambda_i}\left(\psi^\dagger\Phi\right)_i.
\end{equation}
The projections of $\Phi$ onto the eigenstates of $M$ are complex
conjugates, so if the eigenvalues were positive, we would be guaranteed that
$\Phi^\dagger M^{-1}\Phi$ is positive. The trick then is to look
out for an operator $M$ with positive eigenvalues.
In \secref{sec:signProblem} we will see that most Dirac operators $D$ used in
lattice field theory are $\gamma_5$-hermitian\index{hermitian},
which means they obey
\begin{equation}
D^\dagger=\gamma_5 D \gamma_5.
\end{equation}
In that Section, we will show this guarantees
\begin{equation}
\det D = \det D^\dagger,
\end{equation}
and hence if we choose $M=D^\dagger D$, we will have
by \equatref{eq:pseudofermionIdentity}
\begin{equation}\label{eq:twoPseudofermionFlavors}
\left(\det D\right)^2
=\det D^\dagger D
\sim \int\DD{\Phi_R}\DD{\Phi_I}
\exp\left(-\Phi^\dagger \left(D^\dagger D\right)^{-1}\Phi\right),
\end{equation}
i.e. we will be able to rewrite the fermion determinants corresponding
to a system with two identical flavors of quark in a way that guarantees
we can use a Markov chain.
An important practical point to consider is how to generate $\Phi$ according to
the distribution \equatref{eq:twoPseudofermionFlavors}. Let $R$ be a random
vector such that
\begin{equation}
\Phi=D^\dagger R.
\end{equation}
Then if we draw $R$ from a Gaussian, i.e. if we draw
according to $\exp\left(-R^\dagger R\right)$, we are guaranteed that
$\Phi$ will be distributed as desired.
\bibliographystyle{unsrtnat}
\bibliography{bibliography}