-
Notifications
You must be signed in to change notification settings - Fork 0
/
ch_QFT.tex
2023 lines (1875 loc) · 92.2 KB
/
ch_QFT.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
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter{Physics: QFT crash course}
The Standard Model (SM) of particle physics classifies all known
\index{particle!elementary}\index{Standard Model}
{\it elementary particles}, i.e. particles with no known substructure,
and describes three fundamental forces:\index{force!fundamental} the electromagnetic,
weak, and strong forces. Elementary particles can be divided into
\index{particle!matter}
{\it matter particles} (quarks and leptons); {\it gauge bosons}, which mediate
\index{boson!scalar}\index{boson!gauge}
the three aforementioned forces; and a {\it scalar boson}, the Higgs boson,
whose field interacts directly with elementary particles that thereby
acquire their mass. For each particle there exists a corresponding
antiparticle; sometimes a particle is its own antiparticle.
Figure~\ref{fig:SM} gives a schematic overview of the SM.
\begin{figure}
\centering
\includegraphics[width=0.80\linewidth]{figs/SM.png}
\caption{Summary of elementary SM particles. The first three columns give
the three generations of matter particles. Image taken
from the Physics Institute at University of
Zurich~\cite{zurich_SM}.}
\label{fig:SM}
\end{figure}
The theoretical framework underlying the SM is an example of a Quantum
Field Theory (QFT). QFTs are consistent with both quantum mechanics and
relativity. Lattice gauge theories are a kind of QFT; therefore it is
important for the reader to know a little bit about them. A solid understanding
of QFT I think can only be achieved by taking courses along with a significant
amount of self study. In my case, that self study has required multiple years,
but you're a different person, so mileage may vary.
There are a lot
of different resources one can use to learn about QFT; for example when I was a
grad student I used Peskin and Schroeder~\cite{peskin_introduction_1995}
and Srednicki~\cite{srednicki_quantum_2007}. Far and away the most pedagogical
text book I've encountered is by Schwartz~\cite{schwartz_quantum_2014}, and it's
the one I recommend to newcomer in the field.
Nowadays there are also some very high quality lectures on YouTube,
for instance a series by Tong~\cite{tongQFT}, which I found had some other nice
introductory remarks.
A timeline of particle discoveries can be found in
Ref.~\cite{wiki_particle_discoveries}. Another detailed historical overview
of the SM is given in Chapter 1 of Ref.~\cite{griffiths_introduction_2007}.
\section{An mnemonic history of the SM}
One could argue that the beginning of the SM history coincides with the
beginnings of modern particle physics. Since that depends on unifying
relativity, quantum mechanics, and field theory, one could arguably even take
Maxwell's equations as a starting point.
There were also many interesting ideas that were not pursued or turned out
not to be correct yet still played some role in the history; I will not
discuss these. In some cases I may miss some discoveries that were also
important but less celebrated.
Given these ambiguities and the fact
that I am not at all a real historian,
one might call what follows an ``approximate" history.
As I was writing this, I realized that I was trying to tell a story, i.e.
to write it in a way that one development would make sense or feel
motivated given a previous development. Usually that is a bit of an
oversimplification, but it helps me remember why certain discoveries were
significant, where some nomenclature comes from, and what it means. Hopefully it
also helps reveal how physicists think, how we are led to discoveries, and
ultimately why we believe our theories. So with these advantages in mind, I
rather decided to call it a ``mnemonic" history.
Also while I was writing this, I learned a bunch of facts that I found
interesting but are probably a bit off-topic. Hence this mnemonic history is
densely packed with footnotes. For example I decided to start listing Nobel
prizes for some reason. By the time I realized doing this is tedious and does not
teach much, I somehow already felt pot-committed, so I ended up seeing this habit
through to the bitter end.
\subsection{The fundamentals}
% pauli, jordan etc 1920s-30s how to quantize fields.
% Failing to have theory with infinite num of dof,
% then tomonaga, schwinger, dyson, feynman. Can renormalize, QED discovered.
% measurement of Lamb shift 1947. Bethe has explanation that gets refined
% in QED, helps establish QFT.
% 1970s golden age: infinites understood through
% renormalization group wilson and kadanoff.
In 1897 J.J. Thomson did experiments with cathode rays\footnote{In a small
vacuum chamber with two electrodes, if a voltage is applied between them,
electrons will move between them. Televisions used to work by cathode ray tubes,
\index{cathode ray tube} where these electrons are deflected by magnetic fields
to make images on the screen.}
from which he concluded that electric charge must be carried by particles
with high charge-to-mass ratio, the electrons\footnote{He received the 1906
Nobel in physics for this work.}
To explain why atoms are overall electrically neutral, Thomson guessed that
electrons are distributed in a sea of positive charge, which is the
well known {\it plum pudding model}.\index{plum pudding model} This was
disproved by Rutherford in his famous gold foil
experiment~\cite{rutherford_scattering_1911}, in which he discovered
the atomic nucleus. Shortly thereafter, he discovered the
proton~\cite{rutherford_collision_1919}.
Bohr proposed his model~\cite{bohr_constitution_1913}
of hydrogen, supposing it to be made of a proton and an electron, which agreed
well with experiment\footnote{He got the 1922 Nobel for his
contributions understanding atomic structure.}. Extending this theory to
heavier elements by supposing
they are also made of only protons and neutrons however fails, since e.g. helium
is four times as heavy as hydrogen. This difficulty would not be sorted out
until the early 1930s, when Chadwick discovered~\cite{chadwick_possible_1932}
the neutron\footnote{1935 Nobel for him.}.
These early discoveries successfully explained many details of the atom; however
the fact that atomic nuclei are made of particles with only positive or zero
electric charge still required explanation.
Hence for some time, physicists
knew there must be some {\it strong force}\index{force!strong} that opposes
Coulomb repulsion and binds nucleons into nuclei.
Such particles held together by strong interactions are called
{\it hadrons}.\index{hadron} Nowadays we also use the terms {\it meson}
\index{meson} and {\it baryon}\index{baryon} to refer to hadrons made of
two quarks and three quarks, respectively\footnote{This naming scheme
comes from particle weights. At the time, known leptons were light,
baryons were heavy, and mesons were somewhere in the middle. In retrospect it
would have been nicer to name them something like $n$-hadrons, but alas it would
take several decades for us to see that hadrons are made of quarks.}.
One of the earliest, important discoveries of the quantized natures of particle
properties is the celebrated Stern-Gerlach
experiment~\cite{gerlach_experimentelle_1922a,gerlach_magnetische_1922b,gerlach_experimentelle_1922c}.
In this experiment, silver atoms
are deflected by an inhomogeneous magnetic field.
Besides having demonstrated that particles have intrinsic spin, it showed that
the spin is quantized and that measurements of spins along perpendicular axes
``reset" the spin state, and it provided the first measurement of the electron
magnetic moment.
Around this time, physicists were also beginning to see the particle nature of
light. In particular, Planck proposed~\cite{Planck:1901tja}
that light may come in discrete packets of
energy in order to avoid the \index{ultraviolet catastrophe}ultraviolet
catastrophe\footnote{1918 Nobel.}.
Einstein took this proposal seriously~\cite{Einstein:1905cc},
and used it to explain the photoelectric
effect\footnote{1922 Nobel for him. Also in 1905 he published his first
papers on special relativity, as well as a paper on Brownian motion.}.
A careful study~\cite{millikan_direct_1916} of the photoelectric effect by
Millikan showed that
Einstein's interpretation explained the photoelectric effect well\footnote{He
got the 1923 Nobel in part for this reason.}. Finally
Compton showed\footnote{He shared the 1927 Nobel for this.}
that light scattered from a particle shifts by the Compton
wavelength\index{wavelength!Compton}
\begin{equation}
\lambda_c=\frac{\hbar}{2mc},
\end{equation}
where $m$ is the target particle's mass, which one can derive by assuming light
is made of particles with zero rest mass~\cite{Compton:1923zz}.
Altogether these discoveries convinced physicists light behaves as a particle
at short enough length scales, which is the usual photon.\index{photon}
If light is to be quantized, it requires a theory that knows about both quantum
mechanics and special relativity, i.e. it needs QFT.
The standard line of thinking can be cast in this way: One starts with
the Schr\"odinger
equation~\cite{Schrodinger:1926gei,Schrodinger:1926vbi,Schrodinger:1926qnk,Schrodinger:1926xyk}
for a spinless, non-relativistic particle
of mass $m$ in the position basis,
\index{Schr\"odinger equation}
\begin{equation}
i\hbar\partial_t\psi=-\frac{\hbar^2}{2m}\nabla^2\psi.
\end{equation}
If we instead use a relativistic Hamiltonian and square the differential
operators on each side, we get the
{\it Klein-Gordon equation}~\cite{Klein:1926tv,gordon_comptoneffekt_1926}
\index{Klein-Gordon equation}
\begin{equation}
-\hbar^2\partial_t^2\psi=\left(-\hbar^2c^2\nabla^2+m^2c^4\right)\psi.
\end{equation}
While this is at least relativistically sensible, one can show that this
squaring of operators
leads to state normalization being time-dependent, i.e. probability is not
conserved. The situation was finally rescued by Dirac\footnote{Dirac
and Schr\"odinger shared the 1933 Nobel.}, who realized that
one could have a relativistically sensible equation that is first-order
in its operators by introducing some matrices and a spin component
to the wavefunction~\cite{Dirac:1928hu,Dirac:1928ej}. The result is the
{\it Dirac equation}\index{Dirac equation}
\begin{equation}
i\hbar\slashed{\partial}\psi=mc\psi.
\end{equation}
The corresponding Hamiltonian for the Dirac equation is traceless, which
tells you that the energy eigenvalues cancel out, i.e.
it suggests there are states of
negative energy. These negative energy states indicate that the theory
has no ground state. In order to prevent this infinite cascade into increasingly
negative energies, he speculated that these infinitely many states are already
occupied, which is referred to as \index{Dirac sea}the {\it Dirac sea};
the Pauli exclusion principle then prevents this infinite descent.
If an electron in the sea were excited, it would leave behind a vacancy
that would manifest itself as a positively charged particle. This was the
prediction of the existence of the \index{positron}positron, which
was discovered\footnote{1936 Nobel.} in 1932 by Anderson~\cite{Anderson:1933mb}.
Later St\"uckelberg~\cite{Stueckelberg:1941rg} and
Feynman~\cite{feynman_theory_1949} would introduce the modern interpretation
of the positron: rather than being a hole left in the Dirac sea,
the previously negative energy states are to be understood as the
positive energy states of a different particle.
One of the last kinds of fermions needed to complete our particle collection
are the neutrinos. Before 1930, there was a problem with $\beta$-decay
\index{decay!beta}
which is any decay emitting an $e^+$ or $e^-$ from
an atomic nucleus: Energy was not conserved. In particular if one assumes
a general $\beta$-decay process functions like
\begin{equation}
A\to B+e^-,
\end{equation}
one can use conservation of four-momentum to find the electron energy.
The measured energy was found to fluctuate and be smaller than what four-momentum
conservation delivers. Pauli
suggested\footnote{Rather than being documented in a publication, this seems to
come from a letter written by Pauli addressed to a conference in T\"ubingen.
It opens, ``Liebe Radioaktive Damen und Herren".} that this missing energy
lies with an as-yet-undetected, weakly interacting particle, the
electron neutrino. The electron neutrino would not be
discovered\footnote{1995 Nobel.}\index{neutrino!electron}
until the mid 1950s by Cowan and Reines~\cite{Cowan:1956rrn}.
\subsection{Weak and strong forces}
\index{interaction!weak}
In the early 1930s, Fermi published\footnote{Apparently he originally attempted
to publish it in {\it Nature}, but they rejected it because
it because ``it contained speculations too remote from reality to be of interest
to the reader".} his theory of the
\index{decay!beta}$\beta$-decay~\cite{fermi_tentativo_1934}
\begin{equation}
n\to\text{p}+e^-+\bar{\nu}_e.
\end{equation}
He introduced an effective 4-point interaction directly linking the four
particles in the above process.
Shortly thereafter, Yukawa~\cite{yukawa_interaction_1935} put forward that this
interaction should include another field with corresponding quantum that
mediates this interaction\footnote{Nowadays we designate as
{\it Yukawa interaction} any interaction between
Dirac fields and scalar fields of the form\index{interaction!Yukawa}
$g\bar{\psi}\phi\psi$ or $g\bar{\psi}i\gamma_5\phi\psi$.},
sort of like how the photon mediates the
electromagnetic interaction. Another salient point of this paper is
the introduction of the {\it Yukawa potential}\index{potential!Yukawa}
giving the potential of a gauge boson of mass $m$:
\begin{equation}
V(r)=-g^2\frac{e^{-\alpha m r}}{r}.
\end{equation}
Here $g$ is the gauge coupling and $r$ is the interaction range. One sees that
massless gauge bosons have a Coulomb-like potential, while massive ones
are suppressed exponentially\footnote{One can also show that the Fourier
transform of this potential is the propagator, which we will discuss later.},
which gives an explanation why the weak force has a short interaction range.
Besides already hinting massive weak bosons, this paper is considered to be
one of the first theories of the strong force; from this perspective the
proton and neutron exchange massive mesons, which therefore have a limited
interaction range\footnote{1949 Nobel.}.
\begin{figure}
\centering
\includegraphics[width=\linewidth]{figs/Diffusion_Cloud_chamber_explained.jpg}
\caption{Cloud chambers consist of a sealed environment with some vapor
of e.g. alcohol. As a charged particle moves through the vapor, it knocks
electrons off the gas; the resulting ions attract the polar molecules,
which leaves a visible trail for a short time. To identify particles,
you can see e.g. if they were deflected. C. T. R. Wilson is generally
credited as the inventor of cloud chambers, and he shared the 1927 Nobel
in physics for it. They were extremely popular to use in experiment
for finding particles until the later invention of the bubble chamber.
Image taken from Wikipedia~\cite{wiki_cloud}.}
\label{fig:cloud}
\end{figure}
An early experimental search of cosmic ray\footnote{A {\it cosmic ray} is a high
energy proton or atomic nucleus that originates somewhere from space. They were
discovered in the early 1910s by Hess, which got him the 1936 Nobel.}
\index{cosmic ray} measurements using
cloud\index{cloud chamber} chambers (see \figref{fig:cloud})
found the muon~\cite{neddermeyer_note_1937}, which was originally
mistaken\footnote{Indeed the muon and pion masses are pretty close to each
other, sitting at about 106~MeV and 140~MeV, respectively.}
as the meson that Yukawa suggested. An experiment in the late 1940s showed that
the muon does not interact very strongly with atomic
nuclei~\cite{conversi_disintegration_1947}, which rules it out as the strong
force mediator. Thankfully for Yukawa the
pion\footnote{Pions\index{pion}\index{meson!pi@$\pi$} are the lightest
mesons. They are made of pairs of up and down quarks.} was
discovered~\cite{lattes_processes_1947} in 1947\footnote{And got Powell
the 1950 Nobel for it. It is actually a bit puzzling that he is the only
recipient of this prize, most obviously because only three other scientists were
on his team. Furthermore this prize credits him for his ``development of
photographic method for studying nuclear processes", even though this method was
pioneered by other physicists such as Blau and Wambacher.}.
\index{interaction!strong}
In the late 1940s and early 1950s, the {\it kaon} ($K$)~\cite{rochester_evidence_1947}
\index{meson!K@$K$} and {\it lambda} ($\Lambda$)~\cite{hopper_evidence_1950}
\index{baryon!l@$\Lambda$} hadrons were discovered. A kaon
consists of light quark and a strange, while a lambda baryon binds two light
quarks with one from a higher generation. {\it Strangeness}\footnote{We now
\index{strangeness}
identify strangeness $S$ as
$$
S\equiv\#\,\text{anti-strange quarks}-\#\,\text{strange quarks}.
$$}
was originally proposed as a conserved quantity to explain the relatively long
lifetimes of these particles~\cite{pais_remarks_1952,gell-mann_isotopic_1953,
pais_baryon-meson-photon_1953,tadao_charge_1953}.
Ne'eman~\cite{neeman_derivation_1961},
Gell-Mann~\cite{gell-mann_symmetries_1962}, and
Zweig~\cite{zweig_su3_1964} proposed\footnote{Gell-Mann would receive the 1969
Nobel for his contributions to understanding elementary particle
classification.} that these hadrons could be classified
according to the irreducible representations of $\SU(3)$, a viewpoint which
Gell-Mann called the \index{eightfold way}{\it eightfold way}\footnote{This
name is inspired by the eightfold path of Buddhism.}, examples of which are
illustrated graphically in \figref{fig:eightfold}. Gell-Mann
referred to the fundamental units as {\it quarks}\footnote{Gell-Mann borrows
this name from an excerpt of James Joyce's {\it Finnegan's Wake} that begins
``Three quarks for Muster Mark".
Gell-Mann was a bit of a fanciful guy I guess.}.
At first it was not clear that this quark viewpoint was more than a
purely mathematical construction, however deep inelastic
scattering\index{scattering!deep inelastic}
experiments at the Stanford Linear Accelerator (SLAC) showed that
protons are made of smaller particles, and are therefore not
elementary~\cite{bloom_high-energy_1969,breidenbach_observed_1969}.
This alone did not convince the community that quarks were
real\footnote{For a while it was fashionable to refer to rather refer to
nucleon constituents as {\it partons}\index{parton}, a term coined
by Feynman.}, but
subsequent discoveries would solidify the quark model,
for example the 1964 discovery~\cite{barnes_observation_1964}
of the\index{baryon!$\Omega$}
$\Omega$ baryon\footnote{An $\Omega$ baryon is any baryon not
containing $u$ or $d$ quarks. The $\Omega$ with a $t$ is not expected
to exist in the SM because the $t$ lifetime is too short to
interact strongly. The title of the discovery paper refers
to\index{hyperon} {\it hyperons}, which are any baryons with
at least one $s$ quark. Hence $\Omega$ baryons are a type of hyperon.}.
\begin{figure}[t]
\centering
\includegraphics[width=0.48\linewidth]{figs/Meson_octet.png}
\includegraphics[width=0.48\linewidth]{figs/Baryon_decuplet.png}
\caption{{\it Left}: Spin-0 pseudoscalar meson octet. (See
\tabref{tab:discreteSymm} for a definition of
pseudoscalar.) {\it Right}:
Spin-3/2 baryon decuplet. The $s$ represents strangeness,
with all particles in the same horizontal row having the
same strangeness. Electric charge is represented by $q$,
with all particles along the diagonal having the same
electric charge. Images taken from
Wikipedia~\cite{wiki_eightfold}.}
\label{fig:eightfold}
\end{figure}
From here it was shown possible to formulate a QFT for the strong interaction
based on $\SU(3)$~\cite{fritzsch_advantages_1973}, which we
call quantum chromodynamics (QCD).\index{quantum chromodynamics}
The mediators are called
{\it gluons}\index{gluon} with the adjoint representation delivering
eight possible color combinations.
Gross, Wilczek~\cite{gross_d.j._ultraviolet_1973} and
Politzer~\cite{politzer_reliable_1973} demonstrated
{\it asymptotic freedom}\footnote{They got the 2004 Nobel for this.}
\index{asymptotic freedom} in this QFT, i.e. they showed that the
strong coupling decreases with increasing interaction strength, which is
consistent with the fact that one does not observe free quarks\footnote{At
least not at typical temperatures and densities.}.
This theoretical observation is buttressed by strong coupling
expansions in the lattice formulation, introduced by
Wilson~\cite{wilson_confinement_1974}, which show that the potential
energy between two infinitely heavy quarks grows linearly with
increasing separation
(see \secref{sec:hqfe}).
%creutz_monte_1980
%wilson_RG 1 and 2
We round out this section with a short timeline of discoveries of the
remaining QCD particles. In 1974 the discovery of the
$J/\Psi$-meson\index{psion} or {\it psion}\footnote{The $J/\Psi$ consists
of a $\bar{c}c$ pair. This is also sometimes called {\it charmonium}.
\index{charmonium}} at both Brookhaven National Lab (BNL) and
SLAC~\cite{augustin_discovery_1974,aubert_experimental_1974} demonstrated
the existence of the charm quark\footnote{Richter and Ting got the 1974
Nobel prize in physics for this.}, adding further evidence to
the validity of the quark model. The $J/\Psi$ discovery marks the beginning of a
period of rapid discoveries in particle physics sometimes referred to as the
``November Revolution".\index{November Revolution} The existence of the bottom
quark was demonstrated in 1977 at Fermilab~\cite{herb_observation_1977} when the
$Y$-meson\footnote{A $Y$-meson\index{meson!Y} is a $\bar{b}b$ bound state. This is
sometimes called \index{bottomonium}{\it bottomonium}.} was discovered.
In 1979 we found experimental evidence for the
gluon via indirect observations~\cite{barber_discovery_1979} at the
Deutsches Elektronen-Synchrotron (DESY). In part because it is the
heaviest quark, the top quark would not be discovered until
1995~\cite{abachi_observation_1995,abe_observation_1995}
at Fermilab.
\subsection{Unification}
In the mid 1950s, Lee and Yang~\cite{lee_question_1956} suggested possible
experimental tests to search for parity violation in weak interaction
processes\footnote{Lee and Yang won the 1957 Nobel prize for this.}.
Shortly thereafter, Wu et al.~\cite{wu_experimental_1957} demonstrated parity
violation in the $\beta$-decay of \ce{^{60}Co},
a result which was verified by Garwin et al.~\cite{garwin_observations_1957}.
The theory of the weak interaction was extended by Gell-Mann and
Feynman~\cite{feynman_theory_1958} to accommodate parity violation by
introducing vector-axial currents.
That $\beta$-decay proceeds through vector-axial currents was
experimentally verified shortly thereafter~\cite{goldhaber_helicity_1958}.
The unification of the weak and electromagnetic forces began already with
Glashow in 1961~\cite{glashow_partial-symmetries_1961}, where
he puts forward the $\SU(2)\times \U(1)$ symmetry group.
Still, this theory was not known to be renormalizable.
Also the weak interaction is short range, but this suggests that the mediating boson
should be massive according to Yukawa. On the other hand,
massive gauge bosons superficially spoil gauge invariance.
In superconductivity, Ginzburg-Landau theory~\cite{ginzburg_theory_1950}
gives solutions with effective mass. Nambu applied\footnote{2008 Nobel prize.}
this to particle
physics~\cite{nambu_axial_1960,nambu_dynamical_1961,nambu_dynamical_1961-1},
but this implied the existence of Goldstone modes that are not observed.
Higgs~\cite{higgs_broken_1964} and Brout and Englert~\cite{englert_broken_1964}
noticed\footnote{Higgs and Englert received the 2013 Nobel for this.}
that by strategically choosing
the gauge, one can simultaneously
eliminate the Goldstone modes, add a mass term to gauge bosons, and a scalar
boson, the Higgs boson.\index{Higgs!boson}
We will discuss spontaneous symmetry breaking and Goldstone's theorem
\index{spontaneous symmetry breaking} in \secref{sec:ssb}. The Higgs
mechanism is discussed in detail in \apref{ap:spec_higgs}.
The original Higgs-Brout-Englert mechanism was demonstrated only for massive
QED; Kibble extended this idea to non-abelian
groups~\cite{kibble_symmetry_1967}. Weinberg~\cite{weinberg_model_1967}
and Salam~\cite{salam_weak_1968} applied Kibble's results to Glashow's
$\SU(2)\times\U(1)$ idea\footnote{And shared the 1979 Nobel for it.}.
They demonstrated that one can generate masses
for weak gauge bosons along with electrons and muons, while still leaving
neutrinos massless. This approach also predicted neutral weak currents,
which were discovered shortly thereafter by the Gargamelle
experiment~\cite{hasert_observation_1974}. The $W$ and $Z$ bosons would
be discovered at the European Organization for Nuclear Research (CERN)
in the early 1980s~\cite{aubert_ratio_1983,arnison_experimental_1983}.
In 1963, Cabibbo introduced the {\it Cabibbo angle}\index{Cabibbo angle} allowing
for quark mixing in weak interactions~\cite{cabibbo_unitary_1963} to
explain the lifetimes of heavier hadrons. The suppression of flavor changing
neutral currents was explained in the early 1970s through the GIM\index{GIM mechanism}
mechanism~\cite{glashow_weak_1970}, but in order for this mechanism to work,
one needed full doublets of quarks and leptons.
Then Kobayashi and Maskawa predicted the existence of a third
generation~\cite{kobayashi_cp_1973}, since three quark generations are the
minimal amount needed to allow CP violation in the quark sector\footnote{They
shared the 2008 Nobel along with Nambu.}. The full quark mixing matrix
is known as the CKM matrix.\index{CKM matrix} Neutrino mixing is also handled
through a mixing matrix, the so-called PMNS matrix.\index{PMNS matrix}
We discuss neutrino mixing in detail
in \apref{ap:spec_neutrino}.
In the early 1970s, t'Hooft and Veltman
showed\footnote{1999 Nobel prize for them.} these theories are
renormalizable~\cite{t_hooft_regularization_1972}. Together the Higgs mechanism
and renormalizability of the SM allow one to consistently generate gauge boson
masses while ensuring its applicability at all energy scales.
Furthermore CERN's 2012 discovery of the Higgs
boson~\cite{aad_observation_2012,chatrchyan_observation_2012} shows that Higgs mechanism
corresponds to reality, rather than being just a mathematical trick to
consistently approach massive elementary particles.
\section{Introductory remarks about QFT}
Here I just want to list some things that seem to be true about the universe,
and therefore our underlying theory should reflect these things. For example:
\begin{enumerate}
\item Causal influences seem to be {\it local}, i.e. there is no
action-at-a-distance.
\item Elementary particles are completely and perfectly indistinguishable.
\end{enumerate}
One way to make sense of these two points is to assume the existence of
{\it fields},\index{field} math objects whose pre-image is all space-time.
That the field value depends on its space-time coordinate allows it to be local,
and all elementary particles are viewed as excitations of the field. Since all
particles are excitations of the same object, it is therefore unsurprising that
they would be indistinguishable.
Related to point (1) above, and as already mentioned in the introduction, we
would like our theories to have this property:
\begin{enumerate}
\setcounter{enumi}{2}
\item QFT should be consistent with special relativity.
\end{enumerate}
Demand (3) leads in part to the Klein-Gordon and Dirac equations, and from these
we will find that particle number is not conserved.
A fundamental QFT length scale can be heuristically derived from this statement
as follows: Consider an elementary particle in a box of length $L$. By the
uncertainty principle, we have
\begin{equation}
\Delta p\gtrsim \hbar/2L,
\end{equation}
which means according to relativity,
\begin{equation}
\Delta E\gtrsim \hbar c/2L.
\end{equation}
If the energy uncertainty is large enough, i.e. large enough to support a
particle-antiparticle pair, we conclude
\begin{equation}
\Delta E\approx 2mc^2 \gtrsim\hbar c/2L.
\end{equation}
We then introduce the Compton wavelength\footnote{I guess if
one uses $\hbar$ instead of $h$ this is rather the {\it reduced} Compton
wavelength. But I somehow always work using $\hbar$ instead of $h$, so I opted
to abuse this convention a little.}\index{wavelength!Compton}
$\lambda_c=\hbar/mc$. This argument delivers an interpretation for
$\lambda_c$:
\begin{equation}
L \gtrsim \lambda_c/4
\end{equation}
is a distance threshold\footnote{I have also seen $\lambda_c/2$ as the threshold,
which comes when you think the energy uncertainty only has to be large enough to
support a single particle. But in QFT particles are always created from the
vacuum in particle-antiparticle pairs due to conservation laws.} below which
you have to worry about QFT. Below this scale, you are likely to detect
particle-antiparticle pairs of the species you are examining, which you cannot
distinguish, and it becomes difficult to speak a unique particle at all. In that
sense the Compton wavelength gives a characteristic length scale for a
particle. One can compare this with the particle's de Broglie wavelength
\index{wavelength!de Broglie}$\lambda_v=\hbar/mv$, where it behaves in a well-defined way as a wave.
%\section{The principle of stationary action}
%This section follows a fairly well known and delightful lecture by Feynman
%\cite{caltech}.
\index{limit!classical}\index{non-relativistic limit}
\section{The non-relativistic and classical limits}
In this section I briefly give some intuition for how mundane Newtonian
physics can be recovered from the more esoteric relativistic and quantum
theories. Namely \index{limit!non-relativistic} I want to focus on
the following two phrases:
\begin{enumerate}
\item The non-relativistic limit is $c\to\infty$.
\item The classical limit is $\hbar\to0$.
\end{enumerate}
I do not think I have the understanding to prove anything, but at least I
can provide some ideas and examples that can make you believe these
two statements.
The non-relativistic limit is, I think, the easier to understand. When we
learn about relativity, the speed of light $c$ is taken as a ``cosmic speed
limit"; correspondingly, sending $c\to\infty$ lifts the speed limit, and
so perhaps it's not surprising that Galilean physics is recovered.
More explicitly, we can see what happens to the Lorentz factor $\gamma$ and
Einstein velocity addition formula under these limits. For the former
we find
\begin{equation}
\lim_{c\to\infty}\gamma
=\lim_{c\to\infty}\frac{1}{\sqrt{1-v^2/c^2}}=1,
\end{equation}
i.e. there is no longer any time dilation or length contraction.
Meanwhile when $c\gg v$, we find for the velocity addition formula
\begin{equation}
v_1\oplus v_2
=\frac{v_1/c + v_2/c}{1+v_1v_2/c^2}
\approx \frac{v_1}{c} + \frac{v_2}{c},
\end{equation}
i.e. it reduces to Galilean velocity addition.
For the classical limit, one can look at specific, simple examples, such as
the quantum harmonic oscillator. In QM, the energy levels of this system
are given by
\begin{equation}
E_n=\hbar\omega\left(n+\frac{1}{2}\right).
\end{equation}
In the $\hbar\to0$ limit, one therefore sees that the differences in energy
become continuous rather than discrete. More generally one can look at
the Heisenberg uncertainty relation,
\begin{equation}
\Delta p\Delta x\geq \frac{\hbar}{2},
\end{equation}
and when $\hbar=0$, you are once again allowed to know position and
momentum simultaneously.
% TODO: path integral classical limit
%\section{The path integral}
%\section{The Dirac equation}\index{Dirac equation}
% dirac equation
% in order to be lorentz invariant we need gamma_0
% slash notation, maybe algebra rules
% solutions to dirac eq. look like plane waves (3.45)
% helicity as interpretation of spin against momentum
% a dirac spinor can be broken down into into creation/annihilation
% operators over all possible momenta (fourier), sum over all spins,
% attached to plane wave u.
\section{Discrete symmetries}\label{sec:discreteSymm}
A general local meson operator/current/bilinear\footnote{Sometimes in lattice
QCD you see bilinears of the form $\bar{\psi}^f(x)\Gamma\psi^g(x\pm a\muh)$,
which is called a {\it point-split}\index{point-split} operator.} has the form
\begin{equation}\label{eq:mesonInterpLat}
\bar{\psi}^f(x)\Gamma\psi^g(x)
\end{equation}
for fermion flavors $f$ and $g$.
% you can then think about what P will do to these creation/annihilation
% states and the plane waves. has the effect of swapping the top and
% bottom components of u with a flipped spatial momentum, which is
% the same as multiplying by gamma_0.
% this gives you the rule for how psi transforms, from which you can
% figure out everything else
% compare your table with p 71 of P&S
\begin{table}\index{vector}\index{scalar}\index{pseudovector}
\index{pseudoscalar}\index{tensor}
\begin{tabularx}{\linewidth}{LCCr} \hline\hline
Type & $J^{PC}$ & $\Gamma$ & Example particles \\\hline
scalar & $0^{++}$ & $\id$ & $H$\\
pseudoscalar & $0^{-+}$ & $\gamma_5$ & $\pi^0$, $\pi^{\pm}$\\
vector & $1^{--}$ & $\gamma^\mu$ & $\rho^0$, $\rho^{\pm}$\\
pseudovector & $1^{++}$ & $\gamma^\mu\gamma_5$ & $a_1$\\
tensor & $1^{+-}$ & $\sigma^{\mu\nu}$ & $b_1$\\
\hline\hline
\end{tabularx}
\caption{The $J^{PC}$ for various meson currents, where $\Gamma$ indicates the
object sandwiched between the Dirac fields in \equatref{eq:mesonInterp}.
The words ``scalar", ``vector", and ``tensor" indicate how an operator
transforms under Lorentz transformations. The ``pseudo" objects have the
opposite sign changes under parity as their corresponding objects.}
\label{tab:discreteSymm}
\end{table}
\index{meson!rho@$\rho$}\index{meson!pi@$\pi$}\index{Higgs!boson}
For instance the $\rho$ meson is a vector meson\footnote{The $\rho$ meson is any
spin-1 combination of $u$ and $d$
quarks/antiquarks, hence their designation as ``vector" mesons in
\tabref{tab:discreteSymm}. Each quark has spin-1/2, so the $\rho$ meson is the
aligned state, while the pion is the anti-aligned state. The mass of a
$\rho$ meson is much greater than the pion (more than five times greater), which
reflects the fact that under strong interactions, aligned-spin states are higher
energy states.}.
\section{Quantum electrodynamics}\label{sec:qed}
\index{quantum electrodynamics}
\section{Spontaneous symmetry breaking}\label{sec:ssb}
\index{spontaneous symmetry breaking}
We follow Section~11.1 of Peskin and
Schroeder~\cite{peskin_introduction_1995}.
Let $\phi(x)$ denote a vector (in the mathematical sense) of $N$ real,
scalar fields $\phi^i(x)$. Then the Lagrangian
\begin{equation}
\Lagr=\frac{1}{2}\left(\partial_\mu\phi\right)^2
+\frac{1}{2}\mu^2\phi^2-\frac{\lambda}{4}\phi^4
\equiv\frac{1}{2}\left(\partial_\mu\phi\right)^2
+V(\phi)
\end{equation}
is invariant under the orthogonal group\footnote{Recall orthogonal
transformations\index{group!orthogonal} $R$ are the ones with
$R^TR=\id$.} O$(N)$. This is the Lagrangian of the\index{sigma model}
{\it linear sigma model}. Note that is is a generalization of $\phi^4$ theory,
but we have replaced the positive mass parameter $m^2$ with a
negative parameter $-\mu^2$ and rescaled $\lambda$ to eliminate a factor of 6.
Classically, the potential is minimized when $\phi$ lies
on an $N$-dimensional sphere of radius $\sqrt{\mu^2/\lambda}$, i.e.
it is minimized for vectors $\phi_\text{min}$ satisfying
\begin{equation}
\phi_\text{min}^2=\frac{\mu^2}{\lambda}.
\end{equation}
To interpret the theory, we first choose coordinates so that
$\phi_\text{min}$ lies
entirely along the $N$ direction
\begin{equation}\label{eq:phidir}
\phi_\text{min}=(0,0,...,0,v),
\end{equation}
where $v=\sqrt{\mu^2/\lambda}$ is the {\it vacuum expectation value} or VEV.
Then, we define a set of shifted fields $\pi_k$ and $\sigma$ relative
to this point by writing
\begin{equation}\label{eq:phishift}
\phi(x)=\left(\pi_1(x),\pi_2(x),...,\pi_{N-1}(x),v+\sigma(x)\right)
\end{equation}
Written in terms of $\pi$, the $N-1$ dimensional vector with components
$\pi_k$, and $\sigma$, the new Lagrangian becomes
\begin{equation}\label{eq:brokenlagr}
\Lagr=\frac{1}{2}\left(\partial_\mu\pi\right)^2
+\frac{1}{2}\left(\partial_\mu\sigma\right)^2
-\frac{1}{2}\left(2\mu^2\right)\sigma^2-\sqrt{\lambda}\mu\sigma^3
-\sqrt\mu\pi^2\sigma
-\frac{\lambda}{4}\sigma^2
-\frac{\lambda}{2}\pi^2\sigma^2
-\frac{\lambda}{4}\pi^4,
\end{equation}
where we have removed constant terms, because they do not change the
physics. Equation~\eqref{eq:brokenlagr} is the Lagrangian of $N-1$
massless, dynamic fields $\pi_k$ and a dynamic field $\sigma$ with mass
$\sqrt{2}\mu$. Written in this form, the original $\text{O}(N)$ symmetry
is now obscured. There is a remaining $\text{O}(N-1)$ symmetry
rotating the $\pi_k$ among themselves. This is an example of
{\it spontaneous symmetry breaking} (SSB), and we say something like
``the original $\text{O}(N)$ symmetry spontaneously breaks to
the subgroup $\text{O}(N-1)$."
\begin{figure}[t]
\centering
\includegraphics[width=0.8\linewidth]{figs/symm_break.pdf}
\caption{Linear sigma model potential for $N=2$. In (a) the $\phi$ field
mass term $m^2>0$, while (b) gives this potential when
$m^2$ is replaced by a negative parameter $-\mu^2$.
Oscillations along the circle of minima in (b) correspond to
the $\pi$ field. Oscillations in the radial direction correspond
to the $\sigma$ field. Image taken from
Thompson~\cite{thomson_modern_2013}.}
\label{fig:ssb}
\end{figure}
Let's try to gain some geometric intuition for this phenomenon.
Looking at \equatref{eq:phidir}, we see that in $\phi$ space, the $\sigma$
field corresponds to oscillations of $\phi$ orthogonal to the $N-1$
dimensional hypersurface, while the massless $\pi_k$ fields
corresponds to oscillations along the hypersurface.
An example with $N=2$ is shown in \figref{fig:ssb}.
If we take the ground state vector~\eqref{eq:phidir} and hit it with
$\text{O}(N)$, it will be rotated somewhere else on the hypersurface.
The subgroup $\text{O}(N-1)$ hits the first $N-1$ components of the
ground state, which are all 0, thereby leaving it unchanged.
In the original $\phi^4$ theory with
$m^2>0$, the ground state vector was 0, so the $\text{O}(N)$ symmetry
was also a symmetry of the ground state. After SSB, $\text{O}(N)$ changes
the ground state vector in general,
which is why we say the symmetry is broken. Generally any symmetry
respected by the Lagrangian but not by the ground state vector
is a broken symmetry.
In the linear sigma model, massless $\pi$ particles appeared after
SSB. This is a special example of a general result
known as Goldstone's theorem. The generated massless particles are
referred to as \index{Goldstone!boson}{\it Goldstone bosons}.
Many light bosons can be interpreted
as approximate Goldstone bosons; as we will see in \secref{sec:cscont},
the pion can be viewed in this manner.
\begin{theorem}{Goldstone's theorem}{}
\index{Goldstone!theorem}
Consider a Lagrangian of the form
$$
\Lagr=(\text{kinetic term for $\phi$})+
(\text{terms independent of $\phi$})-V(\phi),
$$
where $\phi$ is the $N$-dimensional vector of real, scalar fields
$\phi_k$, and $\Lagr$ is invariant under a continuous, global
transformation of $\phi$ with generators $T^a$. Then for every
spontaneously broken generator there exists a corresponding
Goldstone boson.
\begin{proof}
Let $\phi_\text{min}$ be a constant field minimizing $V$. Expanding
$V$ about this minimum we get to leading order
$$
V(\phi)= V(\phi_\text{min})
+\frac{1}{2}
(\phi-\phi_\text{min})_i(\phi-\phi_\text{min})_j\,
\frac{\partial^2 V}{\partial\phi_i\partial\phi_j}
\Big|_{\phi=\phi_\text{min}}.
$$
The differences $\phi-\phi_\text{min}$ give the new fields of the
theory after SSB; for example in the linear sigma model this difference
is, from \equatref{eq:phidir} and \eqref{eq:phishift},
$$
\phi-\phi_\text{min}=(\pi_1,...,\pi_{N-1},\sigma).
$$
Therefore the coefficient of the quadratic term is a symmetric matrix
whose eigenvalues give the masses of these fields. If we can prove
that each broken generator implies a zero eigenvalue for this matrix,
we are done.
The kinetic term for $\phi$ is already invariant under the global
transformation, so if $\Lagr$ is invariant, it follows that $V$ must
be as well. Then we can write
$$
V\left((\id-i\omega^aT^a)\phi\right)=V(\phi),
$$
where $\omega$ is some infinitesimal parameter. Expanding to linear
order yields
$$
\frac{\partial V}{\partial\phi_j}(T^a\phi)_j=0.
$$
Differentiating the above with respect to $\phi_i$ and evaluating at
$\phi_\text{min}$ gives
$$
\frac{\partial^2 V}{\partial\phi_i\partial\phi_j}
\Big|_{\phi=\phi_\text{min}}(T^a\phi_\text{min})_j=0,
$$
i.e. $T^a\phi_\text{min}$ is annihilated by the mass matrix.
If $T^a$ is a broken generator, we have $T^a\phi_\text{min}\neq0$,
so the above equation implies $T^a\phi_\text{min}$ is an
eigenvector of the mass matrix with eigenvalue zero.
\end{proof}
\end{theorem}
Note that from \tabref{tab:lie}, the difference between the number of generators
of $\ON(N)$ and $\ON(N-1)$ is $N-1$, which is consistent with $N-1$ pions.
\section{Isospin and hypercharge}\label{sec:isohyper}
\index{hypercharge}\index{isospin}
We follow Chapter~9 of
Ref.~\cite{thomson_modern_2013}.
In quantum mechanics you learn that spin is a quantum number of
charged particles. For example the electron is a spin-1/2 particle. The
$z$-component $S_3$ of the spin operator $S$ commutes with the Hamiltonian, and
you learn that the eigenvectors of $S_3$ are the +1/2 and -1/2 states. You also
learn that the components of $S$ are related to the Pauli matrices by
\begin{equation}
S_i=\frac{\hbar}{2}\sigma_i.
\end{equation}
Early on in nuclear physics, scientists noticed that the proton and neutron had
about the same mass, and that the nuclear force between two nucleons (i.e.
protons or neutrons) was approximately charge independent. Therefore Heisenberg
suggested that protons and neutrons were two states of a single particle (the
nucleon) just as there are spin-up and spin-down states of a spin-1/2
particle.
The quantum number corresponding to this property is called {\it isospin}.
Using this idea, the proton and neutron form an isospin doublet with total
isospin $I=1/2$ and $z$-component $I_3=\pm1/2$. Thus the Pauli matrices also
give a suitable representation of the isospin operator, and we write
\begin{equation}
I^2=\sum\limits_{i=1}^3 I_i^2, \qquad I_i=\frac{1}{2}\sigma_i.
\end{equation}
I know this is sloppy, but I want to leave it to the reader to determine from
context whether $I$ represents the operator or the eigenvalue.
The concept of isospin can be extended in the same way to quarks.
In the simplest case we have $N_f=2$ and consider the lightest quarks
$u$ and $d$. The $\SU(2)$ isospin symmetry is only approximate because
the $u$ and $d$ quarks have slightly different masses. The isospin
doublet then has a $u$ component and a $d$ component. Generally with
$N_f$ flavors of fermion with degenerate masses, the symmetry group is
$\SU(N_f)$ and we
form $N_f$-component multiplets in flavor space, one component per flavor.
Let's introduce the $s$ quark, assuming it has the same mass as the
$u$ and $d$ quarks, and do the $N_f=3$ case.
Using the Pauli matrices and the fact that $\SU(3)$ has 8 generators, you can
figure out what the Gell-Mann matrices are. We will say that $u$, $d$, and $s$
are eigenvectors of isospin and write
\begin{equation}
u=\colvec{3}{1}{0}{0}, \qquad
d=\colvec{3}{0}{1}{0}, \qquad
s=\colvec{3}{0}{0}{1}.
\end{equation}
Then $u$ and $d$ span a 2D subspace of flavor space, so from the earlier
discussion we should know that the generators of $\SU(2)$ are contained in the
generators of $\SU(3)$. Hence
\begin{equation}
\lambda_1=\left(\begin{array}{ccc}
0 & 1 & \\
1 & 0 & \\
& & 0
\end{array}\right), \qquad
\lambda_2=\left(\begin{array}{ccc}
0 & -i & \\
i & 0 & \\
& & 0
\end{array}\right), \qquad
\lambda_3=\left(\begin{array}{ccc}
1 & 0 & \\
0 & -1 & \\
& & 0
\end{array}\right).
\end{equation}
But there's nothing special about $u$ and $d$; $u$ and $s$ will similarly form
a subspace, and so will $d$ and $s$. In both cases, we will use the Pauli
matrices as generators. We can similarly write
\begin{equation}
\lambda_4=\left(\begin{array}{ccc}
0 & & 1\\
& 0 & \\
1 & & 0
\end{array}\right), \qquad
\lambda_5=\left(\begin{array}{ccc}
0 & & -i \\
& 0 & \\
i & & 0
\end{array}\right), \qquad
\lambda_X=\left(\begin{array}{ccc}
1 & & 0 \\
& 0 & \\
0 & & -1
\end{array}\right),
\end{equation}
\begin{equation}
\lambda_6=\left(\begin{array}{ccc}
0 & & \\
& 0 & 1\\
& 1 & 0
\end{array}\right), \qquad
\lambda_7=\left(\begin{array}{ccc}
0 & & \\
& 0 & -i\\
& i & 0
\end{array}\right), \qquad
\lambda_Y=\left(\begin{array}{ccc}
0 & & \\
& 1 & 0\\
& 0 & -1
\end{array}\right).
\end{equation}
Finally note the fact that there are only 8 linearly independent generators, so
two of these matrices should be linearly dependent. Since the
$u\leftrightarrow d$ isospin symmetry is the closest to being exact, we choose
the last generator to be a linear combination of $\lambda_X$ and $\lambda_Y$
that treats the $u$ and $d$ quarks symmetrically. Thus
\begin{equation}
\lambda_8=\frac{1}{\sqrt{3}}\lambda_X+\frac{1}{\sqrt{3}}\lambda_Y
=\frac{1}{\sqrt{3}}\left(\begin{array}{ccc}
1 & & \\
& 1 & \\
& & -2
\end{array}\right).
\end{equation}
The new isospin and total isospin operators are
\begin{equation}
I^2=\frac{1}{4}\sum\limits_{i=1}^8 I_i^2, \qquad I_i=\frac{1}{2}\lambda_i.
\end{equation}
In the case of $\SU(2)$, the operators $I_i$ do not commute, and therefore are
not simultaneously diagonalizable. For $\SU(3)$ in the Gell-Mann basis,
$I_3$ and $I_8$ are both diagonal, so they correspond to compatible
observables. The observable $I_3$ is itself sometimes referred to
as isospin or weak isospin. The observable we associate with $I_8$ is rescaled as
\begin{equation}
Y=\frac{1}{\sqrt{3}}\lambda_8,
\end{equation}
and $Y$ is called the {\it hypercharge}.\index{hypercharge}
We can quickly obtain the isopins and hypercharges of
our lightest quarks:
\begin{equation}\begin{aligned}
\hat{I_3}\,u =\frac{1}{2}u ~~~ &\text{and}~~~\hat{Y}u=\frac{1}{3}u,\\
\hat{I_3}\,d =-\frac{1}{2}d ~~~ &\text{and}~~~\hat{Y}d=\frac{1}{3}d,\\
\hat{I_3}\,s = 0 ~~~ &\text{and}~~~\hat{Y}s=-\frac{2}{3}s.
\end{aligned}\end{equation}
One can summarize this pattern of conserved charges in the following formula:
\begin{theorem}{Gell-Mann-Nishijima Formula}{}
\index{Gell-Mann-Nishijima formula}
The electric charge $Q$ of a particle is related to its isospin and
hypercharge by
\begin{equation*}
Q=I_3+\frac{1}{2}Y
\end{equation*}
\end{theorem}
\section{Chiral symmetry}\label{sec:cscont}
\index{chiral symmetry}
There are some interesting, approximate global symmetries of the QCD Lagrangian
that crop up in the limit of zero quark mass. They are related to isospin
discussed in \secref{sec:isohyper}. Introducing small quark masses breaks this
symmetry; hence from the perspective of an effective theory of bound states we
can find something like Goldstone bosons in accordance with the discussion in
\secref{sec:ssb}.
In this section we work with a Euclidean metric. We follow
Chapter 7 of Gattringer and
Lang~\cite{gattringer_quantum_2010}.
The massless fermion action for a single flavor reads
\begin{equation}
S_F=\int\dd[4]{x}\Lagr_F=\int\dd[4]{x}\bar{\psi}\slashed{D}\psi.