-
Notifications
You must be signed in to change notification settings - Fork 5
/
Copy pathlinear-algebra.tex
1444 lines (1144 loc) · 74 KB
/
linear-algebra.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
\documentclass[a4paper]{scrartcl}
\usepackage[
fancytheorems,
fancyproofs,
noindent,
]{adam}
\title{Linear Algebra}
\author{Adam Kelly (\texttt{ak2316@cam.ac.uk})}
\date{\today}
\allowdisplaybreaks
\begin{document}
\maketitle
% This is a short description of the course. It should give a little flavour of what the course is about, and what will be roughly covered in the notes.
This article constitutes my notes for the `Linear Algebra' course, held in Michaelmas 2021 at Cambridge. These notes are \emph{not a transcription of the lectures}, and differ significantly in quite a few areas. Still, all lectured material should be covered. As a warning, this course has lots of results which involve just checking that a definition holds -- such proofs have been intentionally omitted.
\tableofcontents
\section{Vector Spaces}
\subsection{Vector Spaces and Subspaces}
Linear algebra is, somewhat obviously, primarily about studying objects that are \emph{linear} in nature. The objects we really care about are \emph{vector spaces}, settings in which we can add elements and multiply by scalars. We are also going to consider \emph{linear maps}, functions on vector spaces which preserve that linear structure -- but more on that later.
Throughout the following discussion (and this course), $\F$ is going to denote an arbitrary field\footnote{A field $\F$ is a set $\F$ equipped with two operations $+$ (`addition') and $\cdot$ (`multiplication'). We require $\F$ with addition to form an abelian group, and multiplication must be associative and have an identity element $1$. We also require every element except $0$ to have an inverse with respect to multiplication, and multiplication must be distributive over addition.
Informally, you can think of a field as something you can do arithmetic in.}
\begin{definition}[$\F$-Vector Space ]
An \vocab{$\F$-vector space} is an abelian group $(V, +)$ together with a function $\F \times V \rightarrow V$, written $(\lambda, v) \mapsto \lambda v$ such that the following axioms hold:
\begin{enumerate}[label=(\roman*)]
\item \emph{Distributivity in $V$}. $\lambda(v_1 + v_2) = \lambda v_1 + \lambda v_2$,
\item \emph{Distributivity in $\F$}. $(\lambda_1 + \lambda_2)v = \lambda_1 v + \lambda_2 v$,
\item \emph{Associativity}. $\lambda(\mu v) = (\lambda \mu) v$,
\item \emph{Identity}. $1v = v$.
\end{enumerate}
\end{definition}
We usually call elements of $V$ \vocab{vectors} and elements of $\F$ \vocab{scalars}. The identity element in $V$ is usually called the zero vector, and is written $0_V$ (or just $0$ if the context is clear).
If $\F$ is $\R$ or $\C$, we use the terms `real vector space' and `complex vector space', since they're so common.
\begin{example}[Examples of Vector Spaces]~
\vspace{-1.5\baselineskip}
\begin{enumerate}[label=(\roman*)]
\item The set of triples
$$
\{(x, y, z) \mid x, y, z \in \R\}
$$
forms a real vector space called $\R^3$, because you can add any two triples component wise.
\item The set
$$
\Q[\sqrt{2}] = \{a + b\sqrt{2} \mid a, b \in \Q\}
$$
is a $\Q$-vector space, where we add elements and scale by rational numbers in the obvious way.
\item The set $\mathcal{C}[0, 1]$ of all continuous functions $f: [0, 1] \rightarrow \R$ forms a real vector space.
\end{enumerate}
\end{example}
As with many new objects, it's helpful to be able to discuss its substructure. In the case of a vector space $V$, there's a pretty natural notion for what it means for a subset $U \subseteq V$ to still act like a vector space.
\begin{definition}[Subspace]
Let $V$ be a $\F$-vector space.
A subset $U \subseteq V$ is a \vocab{subspace} of $V$ if $U$ is also an $\F$-vector space. If $U$ is a subspace of $V$, we will write $U \leq V$.
% \begin{enumerate}[label=(\roman*)]
% \item $0 \in U$,
% \item $u_1, u_2 \in U$ implies that $u_1 + u_2 \in U$,
% \item $\lambda \in \F$, $u \in U$ implies that $\lambda u \in U$.
% \end{enumerate}
\end{definition}
\begin{example}[Examples of Subspaces]~
\vspace{-1.5\baselineskip}
\begin{enumerate}[label=(\roman*)]
\item The set of vectors $\{(x, y, z) \mid x, y, z \in \R, x + y + z = 0\}$ is a subspace of $\R^3$.
\item The set of polynomials with terms of even degree $\{ a_0 + a_2 x^2 + a_4 x^4 + \cdots + a_{2k} x^{2k} \mid \alpha_{2i} \in \R, k \in \N\}$ is a subspace of $\R[X]$, the vector space of polynomials with coefficients in $\R$.
\end{enumerate}
\end{example}
As you would expect, checking that something is a subspace is usually easier than checking all of the axioms for a vector space. In particular, to check that $U$ is a subspace of an $\F$-vector space $V$, you can just check that the following hold:
\begin{itemize}
\item \emph{Zero vector}\footnote{You may wonder why we need to check this when we already check that we are closed under scaling. To see why, notice that we still have to ensure $U$ is non-empty!}. $0_V \in U$,
\item \emph{Closure under addition}. $u_1, u_2 \in U$ to imply $u_1 + u_2 \in U$,
\item \emph{Closure under scaling}. $\lambda \in \F$ and $u \in U$ to imply $\lambda u \in U$.
\end{itemize}
There are various ways in which we can manipulate subspaces, for example we can take the intersection of two subspaces, and we will get back another subspace.
\begin{proposition}[Intersecting Subspaces]
Let $U, W \leq V$. Then $U \cap W \leq V$.
\end{proposition}
\begin{proof}
Since $U$ and $V$ are both subspaces of $V$, we have $0_V \in U \cap V$, and also since they are both closed under addition and scaling, $u_1, u_2 \in U \cap W$ implies that $u_1 + u_2 \in U \cap W$, and $\lambda \in \F$ implies $\lambda u \in U\cap W$. Thus $U \cap W$ is a subspace of $V$.
\end{proof}
However we can't manipulate subspaces however we want and expect magic. For example, the union of two subspaces is generally \emph{not} a subspace, as it is typically not closed under addition. In fact, the union is only ever a subspace if one of the subspaces is contained in the other.\footnote{There are some more exercises of this flavour on the example sheet.}
We can however try to `complete' the union so that it becomes a subspace.
\begin{definition}[Sum of Subspaces]
Let $V$ be a vector space over $\F$, and let $U, W \leq V$. We define the \vocab{sum} of $U$ and $W$ to be the set
$$
U + W = \{u + w \mid u \in U, w \in W \}.
$$
\end{definition}
This definition immediately forces $U + W \leq V$, and indeed it is the minimal such space (in that any subspace of $V$ containing both $U$ and $W$ must also contain $U + W$).
\subsection{Quotient Spaces}
Since a vector space $V$ forms an abelian group $(V, +)$, we are able to take the quotient by any subspace $U \leq V$.
\begin{definition}[Quotient Space]
Let $V$ an $\F$-vector space, and let $U \leq V$. The \vocab{quotient space} $V/U$ is the abelian group $V/U$ equipped with the scalar multiplication $F \times V/U \rightarrow V/U$ written $(\lambda, v + U) \mapsto \lambda v + U$.
\end{definition}
With this definition, we need to check that this scalar multiplication operation
is well defined. Indeed, if $v_1 + U = v_2 + U$ then
\begin{align*}
v_1 - v_2 &\in U \\
\implies \lambda(v_1 - v_2) &\in U \\
\implies \lambda v_1 + U &= \lambda v_2 + U \in V/U,
\end{align*}
so our operation is indeed well defined.
As you would expect, taking a quotient gives you back a vector space.
\begin{proposition}[Quotient Spaces are Vector Spaces]
$V/U$ is an $\F$-vector space.
\end{proposition}
\begin{proof}[Proof Sketch]
Check definitions (most properties are inherited from $V$ being a vector space).
\end{proof}
% \subsection{Spans, Linear Independence, and Steinitz Exchange Lemma}
\subsection{Basis and Dimension}
You are likely informally familiar with the idea of \emph{dimension}, a measure how much freedom exists in a system. Dimensionality is a rather natural concept with respect to vector spaces, but we will need to move through some technicalities to establish the results we want.
To discuss the amount of freedom, we first need a way to quantify what it means for a set of vectors to be independent from one another. This is the idea of \emph{linear independence}.
\begin{definition}[Linear Independence]
We say that $\{v_1, \dots, v_n\} \in V$ are \vocab{linearly independent} if
$$
\lambda_1 v_1 + \lambda_2 v_2 + \cdots + \lambda_n v_n = 0
$$
implies that $\lambda_1 = \cdots = \lambda_n = 0$.
\end{definition}
\begin{remark}
For an infinite subset $S \subseteq V$, we say it's linearly independent if every finite subset is linearly independent.
\end{remark}
If a set of vectors is \emph{not} linearly independent, then there's some vector in that set that can be written as a linear combination of the others -- so it's not independent of them!
The next idea we need to pin down is being able to see if our set of vectors can `generate' the rest of our vector space.
\begin{definition}[Span]
Let $V$ be a vector space over $\F$, and let $S \subset V$. We define the \vocab{span} of $S$, $\langle S \rangle$ to be the set of finite combinations of elements of $S$.
If $\langle S \rangle = V$, then we say $S$ is \vocab{spans} or \vocab{generates} $V$.
\end{definition}
\begin{remark}
By convention, we also take $\langle \emptyset \rangle = \{ 0\}$. An equivalent definition is that $\langle S \rangle$ is the smallest subspace of $V$ that contains $S$.
\end{remark}
\begin{example}[Quadratic Polynomials]
Let $V$ be the vector space of quadratic polynomials over $\R$,
$$V = \{ax^2 + bx + c \mid a, b, c \in \R\}.$$
Then the subset $S \subseteq V$ with $S = \{1, x, x^2\}$ spans $V$.
\end{example}
% \begin{definition}[Finite Dimension]
% Let $V$ be a vector space over $\F$. We say that $V$ is \vocab{finite dimensional} if it is spanned by a finite set.
% \end{definition}
% \begin{example}[Finite and Infinite Dimensional Vector Spaces]
% The space $\R[X]$ of polynomials with coefficients in $\R$ is infinite dimensional, but the space $\R_n[X]$ of polynomials of degree at most $n$ is finite dimensional, and is spanned by the set $\{1, X, X^2, \dots, X^n\}$.
% \end{example}
% \begin{definition}[Linear Independence]
% We say that $\{v_1, \dots, v_n\} \in V$ are \vocab{linearly independent} if
% $$
% \lambda_1 v_1 + \lambda_2 v_2 + \cdots + \lambda_n v_n = 0
% $$
% implies that $\lambda_1 = \cdots = \lambda_n = 0$.
% \end{definition}
% If a set of vectors is \emph{not} linearly independent, then one of the vectors in the set can be written as a linear combination of the others.
% \begin{remark}
% If $\{v_1, \dots, v_n\}$ are linearly independent, then $v_i \neq 0$ for all $i$.
% \end{remark}
Putting these two concepts together gives us the idea of \emph{bases}, which are sets of linearly independent vectors that span a vector space.
\begin{definition}[Basis]
A subset $S$ of a vector space $V$ is a \vocab{basis} if $S$ is a set of linearly independent vectors that span $V$.
\end{definition}
\begin{example}[Basis for $\R^n$]
The \vocab{canonical basis} of $\R^n$ is the set of vectors
$$
S = \left\{
\begin{pmatrix}1 \\ 0 \\ \vdots \\ 0\end{pmatrix},
\begin{pmatrix}0 \\ 1 \\ \vdots \\ 0\end{pmatrix},
\dots,
\begin{pmatrix}0 \\ 0 \\ \vdots \\ 1\end{pmatrix}
\right\}.
$$
Importantly, this is not the \emph{only} basis of $\R^n$, just one that is quite convenient most of the time.
\end{example}
\begin{remark}
Note that in the definition of a basis there is no requirement for the set of basis vectors $S \subseteq V$ to be finite -- only that any element in $V$ must be representable using finitely many elements of $S$.
\end{remark}
We'd intuitively want to say that the \emph{dimension} of a vector space is the number of elements in its basis. However, we first need to check that this is a well defined notion.
We can at this point distinguish between finite and infinite dimensional vector spaces at this point though\footnote{Can you see why this is well defined already?}.
\begin{definition}[Finite \& Infinite Dimension]
We say a vector space $V$ is \vocab{finite dimensional} if it has a finite basis, and we say it is \vocab{infinite dimensional} otherwise.
\end{definition}
The next result about bases we will prove is that they induce \emph{unique} representations of elements in the vector space.
\begin{lemma}[Unique Representations with a Basis]
Let $V$ be a vector space over $\F$. Then $S \subseteq V$ is a basis of $V$ if and only if any vector $v \in V$ can be written uniquely as a linear combination of elements $v_1, \dots, v_n \in S$.
\end{lemma}
\begin{proof}
Suppose that $S$ was a basis for $V$.
Then if $v \in V$ can't be written as such a linear combination, then $S$ wouldn't not span $V$, contradicting it being a basis.
Also, if $v$ can be written as such a linear combination non-uniquely, then taking
$$
v = \lambda_1 v_1 + \cdots + \lambda_n v_n = \mu_1 v_1 + \cdots + \mu_n v_n.
$$
where $\lambda_i \neq \mu_i$ for at least one value of $i$, we'd have
$0 = v - v = (\lambda_1 - \mu_1) v_1 + \cdots + (\lambda_n - \mu_n)v_n$,
and at least one of these coefficients must be non-zero, contradicting $S$ being linearly independent.
Alternatively, if any element in $V$ \emph{can} be written uniquely, then if $v_1, \dots, v_n \in S$ with $\lambda_1 v_1 + \cdots \lambda_n v_n = 0$ implies that $\lambda_1 = \cdots = \lambda_n = 0$, giving that $S$ must be linearly independent. Since $S$ is also spanning by definition, we see that it therefore must be a basis of $V$.
\end{proof}
With that out of the way, we can prove some results about finite dimensional vector spaces which will help us get towards our definition of dimension.
\begin{lemma}[Spanning Sets Contain a Basis]
Let $V$ be a finite dimensional vector space, and let $S = \{v_1, \dots, v_n\}$ be a set of vectors that spans $V$. Then there is some subset of $S$ that is a basis of $V$.
\end{lemma}
\begin{proof}
If $\{v_1, \dots, v_n\}$ is linearly independent, then we are done. If it's not, then (up to reordering) we have $v_n \in \langle\{v_1, \dots, v_{n - 1}\}\rangle$. But then $\langle\{v_1, \dots, v_n\}\rangle = \langle\{v_1, \dots, v_{n - 1}\}\rangle$, so we can not include $v_n$ in our subset. Not including elements in this way repeatedly, since there is finitely many elements in $S$, we must eventually get a linearly independent set that still spans $V$.
\end{proof}
\begin{theorem}[Steinitz Exchange Lemma]
Let $V$ be a finite dimensional vector space. Then if $\{v_1, \dots, v_m\}$ is a set of linearly independent vectors, and $\{w_1, \dots, w_n\}$ spans $V$, then
\begin{enumerate}[label=(\roman*)]
\item $m \leq n$
\item up to reordering, $\{v_1, \dots, v_m, w_{m + 1}, \dots, v_n\}$ spans $V$.
\end{enumerate}
\end{theorem}
\begin{proof}
We will prove this by induction. Suppose we have replaced $\ell \geq 0$ of the $w_i$, and that
$$
\langle \{ v_1, \dots, v_\ell, w_{\ell + 1}, \dots, w_n\}\rangle = V.
$$
If $m = \ell$, we are done, so assume that $\ell < m$. Then since this set is spanning, we can write $v_{\ell + 1} \in V$ as
$$
v_{\ell+1} = \alpha_1 v_1 + \cdots + \alpha_{\ell} v_{\ell} + \beta_{\ell + 1} w_{\ell + 1} + \cdots + \beta_n w_n.
$$
Since having $\beta_i = 0$ for all $\ell + 1 \leq i \leq n$ would violate linear independence, we can suppose without loss of generality that $\beta_{\ell + 1} \neq 0$. We also note that this implies that $\ell + 1 \leq n$, as otherwise this would not be possible.
Then $w_{\ell + 1} \in \langle \{v_1, \dots, v_{\ell + 1}, w_{\ell + 2}, \dots, w_{n} \} \rangle$, and this set spans $V$.
Repeating this process, we will be done after $m$ steps, and we have also shown (at the final step) that $m \leq n$.
\end{proof}
\begin{corollary}[Dimension]
Let $V$ be a finite dimensional vector space over $\F$. Then any two bases of $V$ have the same number of elements, called the \vocab{dimension} of $V$, denoted $\dim V$ or $\dim_\F V$.
\end{corollary}
\begin{proof}
Immediate by Steinitz exchange lemma.
\end{proof}
So using Steinitz exchange lemma we have finally been able to pin down exactly what is meant by the dimension of a vector space -- it's the size of it's basis.
This should match up to the intuitive idea of `freedom' that you had at the start of this section. Freedom in a vector space comes from varying coefficients, and in a basis we can both freely vary coefficients and also reach any element in a vector space uniquely, so the number of independent parameters really is the size of the basis.
Steinitz also gives us a few useful results for free.
\begin{corollary}
Let $V$ be a vector space over $\F$ with finite dimension $n = \dim V$. Then
\begin{enumerate}[label=(\roman*)]
\item Any independent set of vectors has at most $n$ elements, with equality if and only if it's a basis.
\item Any spanning set has at least $n$ elements, with equality if and only if it's a basis.
\end{enumerate}
\end{corollary}
\begin{proof}
Immediate by Steinitz exchange lemma.
\end{proof}
Working with basis and dimension can make the study of vector spaces much easier. For example, Steinitz allows us nicely extend bases. That is,
\begin{center}
\color{blue}
If we have a vector space $V$, and a subspace $U \leq V$, \\then we can pick a nice basis $\{u_1, \dots, u_\ell\}$ of $U$, and \\extend it to a basis $\{u_1, \dots, u_{\ell}, u_{\ell + 1}, \dots, u_n\}$ of $V$.
\end{center}Working with ideas like this can make many results easier to prove, as we will see in the following propositions.
\begin{proposition}[Dimension of the Sum of Subspaces]
Let $U$, $W$ be subspaces of a vector space $V$. if $U$ and $W$ are finite dimensional, then so is $U + W$, and $\dim U + W = \dim U + \dim W - \dim U \cap W$.
\end{proposition}
\begin{proof}
Pick a basis $\{v_1, \dots, v_a\}$ of $U \cap W$, and extend by Steinitz exchange lemma to a basis $\{v_1, \dots, v_a, u_1, \dots, u_b\}$ of $U$, and to a basis $\{v_1, \dots, v_a, w_1, \dots, w_c\}$ of $W$.
It suffices to prove that $\{v_1, \dots, v_a, u_1, \dots, u_b, w_1, \dots, w_c\}$ is a basis of $U + W$.
Clearly this set of vectors spans $U + W$, so we just need to check that they are linearly independent. Suppose that
$$
\sum_{i = 1}^{a} \alpha_i v_i + \sum_{i = 1}^b \beta_i u_i + \sum_{i = 1}^c \gamma_i w_i = 0.
$$
Rewriting,
\begin{equation}\label{eq:thing}
\sum_{i = 1}^{a} \alpha_i v_i + \sum_{i = 1}^b \beta_i u_i = -\sum_{i = 1}^c \gamma_i w_i, \tag{$\dagger$}
\end{equation}
where the LHS is in $U$ and the RHS is in $W$. This implies that
$
\sum_{i = 1}^c \gamma_i w_i \in U \cap W,
$
and can be written as
$
\sum_{i = 1}^c \gamma_i w_i = \sum_{i = 1}^a \mu_i v_i$,
for some $\mu_i$, and then substituting this back into \eqref{eq:thing},
$$
\sum_{i = 1}^a (\alpha_i + \mu_i) v_i + \sum_{i = 1}^b \beta_i u_i = 0
$$
which forces $\beta_i = 0$. A similar argument also gives $\gamma_i = 0$, which then finally forces $\alpha_i = 0$, since $\{v_1, \dots, v_a\}$ is a basis.
\end{proof}
\begin{proposition}[Dimension of the Quotient Space]
If $V$ is a finite dimensional vector space over $\F$ and $U \subseteq V$, then $U$ and $V / U$ are also finite dimensional, and $\dim V = \dim U + \dim V / U$.
\end{proposition}
\begin{proof}
Let $\{u_1, \dots, u_\ell\}$ be a basis of $U$, and extend it via Steinitz exchange lemma to a basis $\{u_1, \dots, u_\ell, w_{\ell+1}, \dots, w_n\}$ of $V$.
It's easy to see that $\{w_{\ell + 1} + U, \dots, w_{n} + U\}$ is a basis of $V/U$, as it clearly spans and linear independence is inherited from it being a basis of $V$. The result then follows.
\end{proof}
\subsection{Direct Sums}
Previously, we were able to look at the substructure of a vector space by looking at subspaces. Given two subspaces, we were then able to construct their sum, which is the set of all linear combinations of elements in each subspace.
When studying a vector space using its subspaces in this way (considering their sum), it can be useful to impose an \emph{additional} constraint about the way in which the subspaces interact. In particular, it can be useful to impose a uniqueness constraint on the linear combinations that are created.
\begin{definition}[Direct Sum]
Let $V$ be a vector space over $\F$, and let $U, W \leq V$. We say that $V$ is the \vocab{direct sum} of $U$ and $W$, written $V = U \oplus W$ if and only if every element $v \in V$ can be decomposed as
$$
v = u + w
$$
with $u \in U$ and $w \in W$, with this decomposition being unique.
\end{definition}
Of course, we can generalize the notion of a direct sum naturally to the case of multiple subspaces, in the way that you would expect.
\begin{remark}[Warning]
We say that $W$ is \emph{a} direct complement of $U$ in $V$. There is \emph{no uniqueness} of such a complement!
\end{remark}
\begin{lemma}
Let $V$ be a vector space and let $U, W \leq V$.
Then the following are equivalent.
\begin{enumerate}[label=(\roman*)]
\item $V = U \oplus W$
\item $V = U + W$ and $U \cap W = \{0\}$
\item For any basis $B_1$ of $U$ and $B_2$ of $W$, the union $B = B_1 \cup B_2$ is a basis of $V$.
\end{enumerate}
\end{lemma}
\begin{proof}
\emph{(ii) implies (i)}. Let $V = U + W$ with $U \cap W = \{0\}$. Then for all $v \in V$, we can write $v = u + w$ with $u \in U$ and $w \in W$. To see that this is unique, suppose that
$$
v = u + w = u' + w'.
$$
Then $(u - u') = - (w-w')$, and thus they are both in $U \cap W$, but the only element of this is $0$, and thus $u = u'$ and $w = w'$, giving us uniqueness.
\emph{(i) implies (iii)}. Let $B_1$ be a basis of $U$ and $B_2$ be a basis of $W$. Then $B = B_1 \cup B_2$ clearly spans $V$, so we need to check that it's linearly independent. Indeed, if
$$
\sum_{u \in B_1} \lambda_u u + \sum_{w \in B_2} \lambda_w w = 0,
$$
then $V = U \oplus W$ implies that since this is the sum of an element of $U$ and an element of $W$, then each is zero, and linear independence follows from $B_1, B_2$ being sets of linearly independent vectors.
\emph{(iii) implies (ii)}.
Clearly we have $V = U+W$, so we just need to check that $U \cap W = \{0\}$. Let $v \in U \cap W$. Then we can write
$$
v = \sum_{u \in B_1} \lambda_u u = \sum_{w \in B_2} \lambda_w w \implies \sum_{u \in B_1} \lambda_u u - \sum_{w \in B_2} \lambda_w W = 0,
$$
and since $B_1 \cup B_2$ is a basis for $V$, we must have $\lambda_u, \lambda_w = 0$ for all $u, w \in B_1, B_2$, implying that $v = 0$.
\end{proof}
This result extends as you would expect for the case of direct products using multiple subspaces.
\section{Linear Maps}
\subsection{Linear Maps and Isomorphisms}
We are now going to turn our attention to \emph{linear maps}, which are maps that preserve the underlying structure of the vector space that they act on.
\begin{definition}[Linear Map]
Let $V, W$ are vector spaces over $\F$.
A map $\alpha: V \rightarrow W$ is a \vocab{linear map} if for all $\lambda_1, \lambda_2 \in \F$ and $v_1, v_2 \in V$ we have
$$
\alpha(\lambda_1 v_1 + \lambda_2 v_2) = \lambda_1\alpha(v_1) + \lambda_2 \alpha(v_2).
$$
\end{definition}
\begin{example}[Examples of Linear Maps]~
\vspace{-1.5\baselineskip}
\begin{enumerate}[label=(\roman*)]
\item Consider the vector space $\mathcal{C}[0, 1]$ of continuous functions $f:[0, 1] \rightarrow \R$, Then for any fixed point $x \in [0, 1]$, the map $\alpha: C[0, 1] \rightarrow \R$ with $\alpha(f) = f(x)$ is a linear map.
\item The map $\R^3 \rightarrow \R$ given by $(a, b, c) \mapsto 4a + 2b + c$ is a linear map.
\item The map $\alpha: Q[\sqrt{2}] \rightarrow Q[\sqrt{2}]$ of `multiplying by $\sqrt{2}$' is a linear map, with $a + b \sqrt{2} \mapsto 2b + a \sqrt{2}$.
\end{enumerate}
\end{example}
It's easy to see that the identity map is a linear map, and also that linearity is preserved by composing linear maps.
One nice thing about linear maps is that they interact well with basis, because of linearity. Indeed, specifying a linear map on the basis is sufficient to define that map on all of a vector space.
\begin{lemma}[Extension by Linearity]\label{lemma:extend-map}
Let $V$ and $W$ be vector spaces, and let $B$ be a basis for $V$. Then if $\alpha_0: B \rightarrow W$ is \emph{any} map, then there exists a \emph{unique} linear map $\alpha: V \rightarrow W$ extending $\alpha_0$, so that for all $v \in B$ we have $\alpha(v) = \alpha_0(v)$.
\end{lemma}
\begin{proof}
For $v \in V$, let $v = \sum_{i = 1}^n \lambda_i v_i$. Then necessarily by linearity, we define
$$
\alpha(v) = \alpha\left(\sum_{i = 1}^n \lambda_i v_i\right) = \sum_{i = 1}^n \lambda_i \alpha_0 (v_i).
$$
By construction this is a linear map and it extends $\alpha_0$.
\end{proof}
As with many other areas of pure mathematics, we have a special case for the structure preserving bijections on a given object.
\begin{definition}[Isomorphism]
If $V$ and $W$ are vector spaces over $\F$, the linear map $\alpha: V \rightarrow W$ is an \vocab{isomorphism} if it also bijective. We say two vector spaces $V$ and $W$ are \vocab{isomorphic}, written $V \cong W$, if there is an isomorphism between them.
\end{definition}
\begin{remark}
if $\alpha:V \rightarrow W$ is an isomorphism, then the inverse map $\alpha^{-1}$ exists and importantly it's \emph{also} linear.
\end{remark}
\begin{lemma}[Isomorphism is an Equivalence Relation]
$\cong$ is an equivalence relation on the class of all vector spaces over $\F$.
\end{lemma}
\begin{proof}
For reflexivity, we note that the identity map $id: V \rightarrow V$ is an isomorphism. For symmetry, we note that if $\alpha: V \rightarrow W$ is an isomorphism, then $\alpha^{-1}: W \rightarrow V$ is also an isomorphism. Lastly, we get transitivity because if $\alpha: U \rightarrow V$ and $\beta: V \rightarrow W$ are isomorphisms, then $\beta \circ \alpha: U \rightarrow W$ is also an isomorphism.
\end{proof}
From one perspective, isomorphism allows us to know when two objects are really just different representations of the same thing. With this in mind, the following result tells us something interesting (yet unsurprising) about the nature of finite dimensional vector spaces.
\begin{theorem}[Isomorphism for Finite Dimensional Vector Spaces]
If $V$ is an $n$ dimensional vector space over $\F$, then $V \cong \F^n$, the vector space
$$
\F^n = \left\{\begin{pmatrix}x_1 \\ x_2 \\ \vdots \\ x_n\end{pmatrix} \mid x_1, \dots, x_n \in \F \right\},
$$
where addition is component-wise.
\end{theorem}
\begin{proof}
Let $B = \{v_1, \dots, v_n\}$ be a basis for $V$. Then we claim that the map $\alpha: V \rightarrow \F^n$ defined by
$$
v = \sum_{i = 1}^n \lambda_i v_i \longmapsto \begin{pmatrix}\lambda_1 \\ \vdots \\ \lambda_n\end{pmatrix}
$$
is an isomorphism. Indeed, it's easy to see that we have linearity inherited from $\F$, and also the map is a bijection, so indeed we do have $V \cong \F^n$.
\end{proof}
In particular, this result tells us that when we choose a basis for a finite dimensional vector space $V$, it's really the same as picking some isomorphism from $V$ to $\F^n$.
This theorem also directly implies (with isomorphism being an equivalence relation) a criterion for two finite dimensional vector spaces to be isomorphic.
\begin{theorem}[Isomorphism Criterion]
Let $V$ and $W$ be finite dimensional vector spaces over $\F$. Then they are isomorphic if and only if they have the same dimension.
\end{theorem}
\subsection{Images and Kernels}
Stepping back from isomorphisms, we need to define some frequently used concepts relating to linear maps.
\begin{definition}
Let $V, W$ be vector spaces over $F$, and let $\alpha: V \rightarrow W$ be a linear map. We define the \vocab{kernel} of $\alpha$ as
$$
\kernel \alpha = \{ v\in V \mid \alpha(v) = 0\},
$$
and the \vocab{image} of $\alpha$ as
$$
\image \alpha = \{ w \in W \mid w = \alpha(v), v \in V\}.
$$
\end{definition}
The structure of a linear map carries over to the image and the kernel, and naturally implies that they are both subspaces of their parent vector spaces.
\begin{lemma}[Image and Kernel are Subspaces]
Let $\alpha: V \rightarrow W$ be a linear map. Then $\kernel \alpha$ and $\image \alpha$ are subspaces of $V$ and $W$ respectively.
\end{lemma}
\begin{proof}[Proof Sketch]
Check definitions.
\end{proof}
You may recognize the definitions of the kernel and image from a course in group theory, particularly in relation to the first isomorphism theorem\footnote{Recall that if $G$ and $H$ are groups, and $\phi: G \rightarrow H$ is a homomorphism, then $G/\kernel{\phi} \cong \image \phi$.}. Indeed, remembering that vector spaces with vector addition are groups, with we can use the kernel and image of a linear map to establish an isomorphism between vector spaces.
\begin{theorem}[First Isomorphism Theorem For Vector Spaces]
Let $V$ and $W$ be $\F$ vector spaces, and let $\alpha: V \rightarrow W$ be a linear map. Then $V/\kernel \alpha \cong \image \alpha$.
\end{theorem}
\begin{proof}[Proof Sketch]
Consider the isomorphism $\overline{\alpha} : V/\kernel \alpha \rightarrow \image \alpha$ defined by $\overline{\alpha}(v + \kernel \alpha) = \alpha(v)$, and check definitions.
\end{proof}
Now while the kernel and image being subspaces are rather straightforward, slightly more interesting is the \emph{dimension} of these subspaces. This is described by the \emph{rank-nullity theorem}.
\begin{definition}[Rank and Nullity]
Let $\alpha$ be a linear map. Then the \vocab{rank} of $\alpha$, $r(\alpha) = \dim \image \alpha$, is the dimension of the image. The \vocab{nullity} of $\alpha$, $n(\alpha) = \dim \kernel \alpha$, is the dimension of the Kernel.
\end{definition}
\begin{theorem}[Rank-Nullity Theorem]
Let $U$ and $V$ be vector spaces such that $U$ is finite dimensional.
Let $\alpha: U \rightarrow V$ be a linear map. Then
$$
\dim U = r(\alpha) + n(\alpha).
$$
\end{theorem}
\begin{proof}
We know that $U/\kernel \alpha \cong \image \alpha$. Thus $\dim (U / \kernel \alpha) = \dim (\image \alpha)$, that is $\dim U - \dim \kernel \alpha = \dim \image \alpha$, or $\dim U = r(\alpha) + n(\alpha)$.
\end{proof}
This result has an important consequence which is incredibly useful for quickly checking if a map is an isomorphism.
\begin{lemma}[Characterisation of Isomorphism]
Let $V$ and $W$ be vector spaces over $\F$ with \emph{equal finite dimension}. Then the following are equivalent:
\begin{enumerate}[label=(\roman*)]
\item $\alpha$ is injective,
\item $\alpha$ is surjective,
\item $\alpha$ is an isomorphism.
\end{enumerate}
\end{lemma}
\begin{proof}[Proof Sketch]
Follows directly from the rank-nullity theorem.
\end{proof}
\subsection{Matrices}
When dealing with vector spaces, it's sometimes convenient to work with vector spaces in the abstract, and other times it's convenient to pick some nice basis which you can easily work with.
When dealing with linear maps $\alpha: V \rightarrow W$, it it can also be useful to work with bases. The way we do this is by introducing \emph{matrices}.
It's important to say that matrices are not the essence of linear algebra, and in may cases viewing the subject in this way will just give you the impression that linear algebra is an opaque subject. Instead, you should think of matrices as a nice representations of linear maps for which its easy to perform computations with.
With that said, let's talk about linear maps.
In this section, we are going to discuss linear maps between two vector spaces $V$ and $W$.
\begin{definition}[Space of Linear Maps]
Let $V$ and $W$ be vector spaces over $\F$. We define
$$
\mathcal{L}(v, w) = \{\alpha: V \rightarrow W \mid \alpha \text{ is a linear map}\}.
$$
\end{definition}
\begin{proposition}[Space of Linear Maps is a Vector Space]
$\mathcal{L}(V, W)$ is a vector space over $\F$. Moreover, if $V$ and $W$ are finite dimensional, then
$$
\dim \mathcal{L}(V, W) = \dim V \cdot \dim W.
$$
\end{proposition}
\begin{proof}[Proof Sketch]
To see that $\mathcal{L}(V, W)$ is a vector space, one just needs to check definitions. We will leave the proof of the dimension until later.
\end{proof}
Now, given some linear map $\alpha$, let's think about how we can represent this map with respect to a basis.
Suppose that we had a finite dimensional vector space $V$ with basis $\{v_1, \dots, v_m\}$, and a finite dimensional vector space $W$ with basis $\{w_1, \dots, w_n\}$, and suppose that we had some linear map $\alpha : V \rightarrow W$.
Given some $v \in V$, if we wanted to work out $\alpha(v)$, we can write express $v$ in the basis above, with
$$
\alpha(v) = \alpha\left(\sum_{i = 1}^m \lambda_i v_i\right) = \sum_{i = 1}^m \lambda_i \alpha(v_i).
$$
So, to compute what $\alpha(v)$ is for any $v \in V$, it suffices to know what $\alpha(v_1), \dots, \alpha(v_m)$ is. To specify what say $\alpha(v_1)$ is, we can use the above basis of $W$ and write
$$
\alpha(v_1) = a_{11} w_1 + a_{21} w_2 + \cdots + a_{n1} w_n,
$$
for some coefficients $a_{i1}$. Repeating this for $v_2, \dots, v_n$ then totally specifies the linear map $\alpha$.
These collection of coefficients can be put into a \emph{matrix}, which will then be a representation of $\alpha$ when coupled with the bases $\{v_i\}$ and $\{w_i\}$.
\begin{definition}[Matrix]
An $n \times m$ \vocab{matrix} over $\F$ is an array with $n$ rows and $m$ columns, with entries in $\F$:
$$
A = \begin{pmatrix}
a_{11} & a_{12} & \cdots & a_{1m} \\
a_{21} & a_{22} & \cdots & a_{2m} \\
\vdots&\vdots&\ddots &\vdots\\
a_{n1} & a_{n2} & \cdots & a_{nm}
\end{pmatrix}.
$$
We denote the set of $n \times m$ matrices over $\F$ by $\mathcal{M}_{n,m}(\F)$.
\end{definition}
\begin{proposition}[Vector Space of Matrices]
$\mathcal{M}_{n, m}(\F)$ is an $\F$-vector space.
\end{proposition}
\begin{proof}[Proof Sketch]
Taking addition entry-wise we can just check definitions.
\end{proof}
\begin{proposition}[Dimension of $\mathcal{M}_{n,m}(\F)$]
$\dim \mathcal{M}_{m, n}(\F) = m \times n$.
\end{proposition}
\begin{proof}
We define the \emph{elementary matrix} $E_{i, j}$ such that every entry is $0$ except for the $i, j$th entry, which is $1$.
Then the set $\{E_{i, j} \mid 1 \leq i \leq m, 1 \leq j \leq n\}$ is clearly a basis, and has $n \times m$ elements.
\end{proof}
Following on from the discussion above, we can now write down how linear maps can be represented using matrices.
\begin{definition}[Matrix of $\alpha$ in the $B, C$ Basis]
Let $\alpha: V \rightarrow W$ be a linear map, and let $B = \{v_1, \dots, v_m\}$ be a basis of $V$ and $C = \{w_1, \dots, w_n\}$ be a basis of $W$. Then we define the \vocab{matrix of $\alpha$} to be
$$
[\alpha]_{B, C} = \begin{pmatrix}
| & | & & | \\
[\alpha(v_1)]_C & [\alpha(v_2)]_C & \cdots & [\alpha(v_m)]_C \\
| & | & & |
\end{pmatrix} = \begin{pmatrix}
a_{11} & a_{12} & \cdots & a_{1m} \\
a_{21} & a_{22} & \cdots & a_{2m} \\
\vdots&\vdots&\ddots &\vdots\\
a_{n1} & a_{n2} & \cdots & a_{nm}
\end{pmatrix},
$$
where $a_{ij}$ is the coefficient of $w_i$ in $\alpha(v_j)$.
\end{definition}
\begin{remark}[Notation]
We use $[v]_C$ to denote the column vector representing the vector $v$ in the basis $C$.
\end{remark}
\begin{proposition}[Matrix Representation of Linear Maps]
If $V$ and $W$ are vector spaces over $\F$ such that $\dim V = n$ and $\dim W = m$, then $\mathcal{L}(V, W) \cong \mathcal{M}_{m, n}(\F)$.
\end{proposition}
\begin{proof}
Fix a basis $B, C$ of $V$ and $W$ respectively.
Then consider the linear map
$
\theta: \mathcal{L}(V, W) \rightarrow \mathcal{M}_{m, n}(\F)
$ given by $\alpha \mapsto [\alpha]_{B, C}$.
This map is clearly linear, and it's also clearly a bijection thus it's an isomorphism and our result follows.
\end{proof}
\begin{remark}
This implies that $\dim \mathcal{L}(L, W) = m \times n$.
\end{remark}
What's nice about matrices is that they make working with linear maps \emph{easy}. It's generally much easier to manipulate arbitrary matrices than it is to arbitrary linear maps (of course, this varies from case to case), and we will see this in the following set of computational results.
You should be familiar from previous courses with the notion of matrix multiplication.
\begin{definition}[Matrix Multiplication]
Let $A = (a_{ij})$ be an $n \times p$ matrix, and $B = (b_{ij})$
be an $p \times m$ matrix.
Then their \vocab{product} $C = A\cdot B$ is an $n \times m$ matrix $C = (c_{ij})$, with
$$
c_{ij} = \sum_{k = 1}^p a_{ik}b_{kj}.
$$
\end{definition}
We can use matrix multiplication in lots of nice ways.
\begin{lemma}[Evaluating Linear Maps]
Let $\alpha: V \rightarrow W$ be a linear map, and let $B, C$ be bases for $V$ and $W$ respectively. Then if $v \in V$, we have
$$
[\alpha(v)]_C = [\alpha_{B,C}] \cdot [v]_B.
$$
\end{lemma}
\begin{proof}
Let $v \in V$, and let $v = \sum_{j = 1}^n \lambda_j v_j$. Then
$$
\alpha(v) = \sum_{j = 1}^n \lambda_j \alpha(v_j) = \sum_{j = 1}^n \lambda_j \sum_{i = 1}^m a_{ij} w_i = \sum_{i = 1}^m \left(\sum_{j = 1}^n a_{ij} \lambda_j\right) w_i,
$$
as required.
\end{proof}
\begin{lemma}[Composing Linear Maps]
Let $\alpha: V \rightarrow W$ and $\beta: U \rightarrow V$ be linear maps. Then if $A, B, C$ are bases of $U, V$ and $W$ respectively, we have
$$
[\alpha \circ \beta]_{A, C} = [\alpha]_{B, C} \cdot [\beta]_{A, B}.
$$
\end{lemma}
\begin{proof}[Proof Sketch]
Use the definition of matrix multiplication to check that this map corresponds to the correct thing.
\end{proof}
% \begin{example}
% Let $V$ and $W$ be finite dimensional vector spaces.
% Consider the linear map $\alpha: V \rightarrow W$, and suppose $Y \leq V$. Consider the image of $Y$ under $\alpha$, $\alpha(Y) = \{ \omega \in W \mid \omega = \alpha(y), y \in Y\}$.
% Define the sets $B' = \{v_1, \dots, v_k\}$ and $B'' = \{v_{k + 1}, \dots, v_n\}$ such that the union of disjoint sets $B = B' \cup B''$ is a basis for $V$, and $B'$ is a basis of $Y$ (that is, we extend using Steinitz exchange lemma).
% Similarity we define $C' = \{w_1, \dots, w_\ell\}$ and $C'' = \{w_{\ell + 1}, \dots, w_m\}$ such that the union of disjoint sets $C = C' \cup C''$ is a basis for $W$ and $C'$ is a basis for $\alpha(Y)$.
% Then we have the matrix
% $$
% [\alpha]_{B, C} = \begin{pmatrix}
% \alpha(v_1) & \cdots & \alpha(v_k) & \alpha(v_{k + 1}) & \cdots & \alpha(v_n)
% \end{pmatrix},
% $$
% and this will look like the block matrix
% $$
% \left[\begin{array}{ c | c }
% A & B \\
% \hline
% 0 & C
% \end{array}\right],
% $$
% where $A$ is $k \times \ell$.
% \end{example}
\subsection{Change of Basis and Equivalent Matrices}
Matrices represent a linear map with respect to a given basis, so a natural question is how do we rewrite a matrix in a different basis? That is,
given two vector spaces $V, W$ and a linear map $\alpha: V \rightarrow W$, we want to take the matrix $[\alpha]_{B, C}$ and express it with respect to the basis $B', C'$.
The way we do this is by introducing a \emph{change of basis} matrix.
\begin{definition}[Change of Basis Matrix]
The \vocab{change of basis} matrix from $B'$ to $B$ is the matrix
$P = [I]_{B', B}$, where $I$ is the identity map.
\end{definition}
\begin{remark}
$P$ is an invertible square matrix, and $P^{-1}$ is the change of basis matrix from $B$ to $B'$.
\end{remark}
Given a vector represented in a basis $B$, the change of basis matrix allows that vector to be expressed in a different basis $B'$.
\begin{lemma}[Change of Basis for a Vector]
Let $B$ and $B'$ be bases for a vector space $V$, and let $v \in V$. Then if $P$ is the change of basis matrix from $B'$ to $B$ we have
$$
[v]_{B} = P[v]_{B'}.
$$
\end{lemma}
\begin{proof}
We know that $P = [I]_{B', B}$, and so $[I(v)]_B = [I]_{B', B} [v]_B' = P[v]_{B'}$.
\end{proof}
Using this idea of change of basis matrices, we can express the matrix of a linear map with respect to another basis.
\begin{proposition}[Change of Basis Formula]
Let $A = [\alpha]_{B, C}$ and $A' = [\alpha]_{B', C'}$, and let $P$ be the change of basis from $B'$ to $B$ and $Q$ be the change of basis matrix from $C'$ to $C$. Then
$$
A' = Q^{-1} AP.
$$
\end{proposition}
\begin{proof}[Proof Sketch]
Check definitions.
\end{proof}
\subsection{Equivalent Matrices}
A natural notion of equivalence comes from saying
two matrices are equivalent if there is some choice of basis for which they represent the same linear map. Using the change of basis formula, we can pin down this notion exactly.
\begin{definition}[Equivalent Matrices]
We say that two matrices $A, A' \in \mathcal{M}_{m, n}(\FF)$ are \vocab{equivalent} if $A' = Q^{-1}AP$ for some invertible matrices $Q \in \mathcal{M}_{m, m}(\F)$ and $P \in \mathcal{M}_{n, n}$.
\end{definition}
It's easy to check that this definition gives us an equivalence relation, and indeed it turns out that the equivalence classes of matrices are remarkably straightforward.
\begin{proposition}[Natural Basis for Matrices]
Let $V, W$ be vector spaces over $\F$, with $\dim V = n$ and $\dim W = m$. Let $\alpha: V \rightarrow W$ be a linear map, Then there exists a basis $B$ of $V$ and a basis $C$ of $W$ such that
$$
[\alpha]_{B, C} = \left(\begin{array}{ c | c }
I_r & 0 \\
\hline
0 & 0
\end{array}\right),
$$
where $r = r(\alpha)$ and $I_r$ is the $r \times r$ identity matrix.
\end{proposition}
\begin{proof}
We first choose a basis $\{v_{r + 1}, \dots, v_n\}$ of $\kernel \alpha$, and extend it to a basis $B = \{v_1, \dots, v_n\}$ of $V$.
We will first show that $\{\alpha(v_1), \dots, \alpha(v_r)\}$ is a basis of $\image \alpha$.
To see that it's spanning, given $v \in V$ we can write
$\alpha(v) = \alpha\left(\sum_{i = 1}^n \lambda_i v_i\right) = \sum_{i = 1}^n \lambda_i \alpha(v_i) = \sum_{i =1}^r \lambda_i \alpha(v_i),
$
since $\alpha(v_i) = 0$ for $i > r$.
To see that we have linear independence, suppose that $\sum_{i = 1}^r \lambda_i \alpha(v_i) = 0$. Then we have
$
\alpha\left(\sum_{i = 1}^r \lambda_i v_i\right) = 0$ implying that
$\sum_{i = 1}^r \lambda_i v_i \in \kernel \alpha$,
and since the kernel is spanned by $\{v_{r + 1}, \dots, v_n\}$ and $\{v_1, \dots, v_n\}$ is a basis, we must have $\lambda_i = 0$, as required.
We can now extend this basis to a basis $C = \{\alpha(v_1), \dots$, $\alpha(v_r), w_{r + 1}, \dots, w_m \}$ of $W$. Then we can express $\alpha$ as a matrix in the basis $B$ and $C$ as
\begin{align*}
[\alpha]_{B, C} &=
\begin{pmatrix}
| & & | & | & & |\\
[\alpha(v_1)]_C & \cdots & [\alpha(v_r)]_C & [\alpha(v_{r + 1})]_C & \cdots & [\alpha(v_n)]_C \\
| & & | & | & & |
\end{pmatrix} \\
&= \begin{pmatrix}
| & & | & | & & |\\
[\alpha(v_1)]_C & \cdots & [\alpha(v_r)]_C & [0]_C & \cdots & [0]_C \\
| & & | & | & & |
\end{pmatrix},
\end{align*}
and it can easily be seen that this matrix is of the form desired.
\end{proof}
This result gives us a clear criterion for the equivalence of matrices, and also gives us a lovely result which is useful for choosing a basis which works nicely with the matrix representation of a linear map.
\begin{corollary}[Matrix Equivalence Criterion]
Two $m \times n$ matrices are equivalent if the linear maps they represent have the same rank.
\end{corollary}
The equivalence criterion above is great, but doesn't give us a very easy way to see what the rank of the underlying linear map is, without doing a large amount of work. What we really want is a way to get this rank directly from the matrix. We can do this by considering the \emph{column} and \emph{row rank} of a matrix (which as we will see are really the same notion).
\begin{definition}[Column and Row Rank]
Let $A \in \mathcal{M}_{m, n}(\F)$. The \vocab{column rank} of $A$, $r(A)$, is the dimension of the subspace of $\F^m$ spanned by the column vectors of $A$.
Similarity, the \vocab{row rank} is the dimension of the subspace spanned by the row vectors of $A$ (that is, it's the column rank of $A^T$).
\end{definition}
\begin{remark}
If $\alpha$ is a linear map, represented by a matrix $A$ with respect to some basis, then $r(A) = r(\alpha)$, so column rank is the same as the rank.
\end{remark}
\begin{proposition}[Equivalence from Row Rank]
Two matrices $A, A'$ are equivalent if and only if $r(A) = r(A')$.
\end{proposition}
\begin{proof}
If they are equivalent, then they correspond to the same linear map $\alpha$, expressed in two different bases. So the column rank is $r(A) = r(A') = r(\alpha)$, and we are done.
If $r(A) = r(A') = r$, then $A$ and $A'$ are both equivalent to
$$
\left(\begin{array}{ c | c }
I_r & 0 \\
\hline
0 & 0
\end{array}\right),
$$
and by transitivity of equivalent, we have $A$ and $A'$ equivalent.
\end{proof}
The distinction between column rank and row rank is really non-existent, as the following result shows. This also means that we can refer to the column and row rank as just the `rank' of the matrix.
\begin{theorem}[Column Rank Equals Row Rank]
$r(A) = r(A^T)$
\end{theorem}
\begin{proof}
Let $r = r(A)$. Then we know that
we can write
$$
Q^{-1}AP = \left(\begin{array}{ c | c }
I_r & 0 \\
\hline
0 & 0
\end{array}\right),
$$
then taking the transpose, we have $(Q^{-1}AP)^T = P^T A^T (Q^{-1})^T = P^T A^T (Q^T)^{-1}$
so $P^T A^T (Q^T)^{-1}$ is
$$
\left(\begin{array}{ c | c }
I_r & 0 \\
\hline
0 & 0
\end{array}\right)^T,
$$
which is an $n \times m$ matrix. But then $r = r(A) = r(A^T)$.
\end{proof}
Let's now consider the special case of \vocab{endomorhisms}, that is, linear maps $\alpha: V \rightarrow V$.
If $B$ and $B'$ are a basis for $V$, and $P$ is the change of basis matrix from $B'$ to $B$, then given $\alpha$ we have $[\alpha]_{B', B'} = P^{-1}[\alpha]_{B, B}P$, which is a special case of the change of basis formula.
This idea of writing a matrix using the same basis for the domain and range gives us \emph{similarity}, another notion of matrices being the same.
\begin{definition}[Similar Matrices]
Let $A, A'$ be two $n \times n$ square matrices. We say that $A$ and $A'$ are \vocab{similar} or \vocab{conjugate} if and only if $A' = P^{-1} A P$ for some invertible $n \times n$ matrix $P$.
\end{definition}
This idea of similarity will be a central concept when we deal with the diagonalization of such matrices later on.
\section{Dual Space}
\subsection{Defining the Dual Space}
We are now going to turn our attention to an interesting set of linear maps on a vector space.
\begin{definition}[Dual Space]
Let $V$ be a vector space over a field $\F$. The \vocab{dual space} $V^*$ is the set of linear maps from $V$ to $\F$,
$$
V^* = \mathcal{L}(V, \F).
$$
\end{definition}
So how should we understand this space?
As usual, things are easier for finite dimensional vector spaces, so we are going to start there. We know from the previous section that \emph{any} linear map is determined by its action on basis vectors. So if we have some finite dimensional vector space $V$ with basis $\{v_1, \dots, v_n\}$ and a linear map $\alpha: V \rightarrow \F$, we just need to know $\alpha(v_1), \dots, \alpha(v_n) \in \F$.
This tells us exactly what the nice way to think about the dual space is -- through a basis that just picks out basis vectors so that we can deal with them individually. We call this the \emph{dual basis}.\footnote{We just write it as a definition here, but we should really check that it is indeed a basis of the dual, but I think it's relatively clear so this is left to the reader.}
\begin{definition}[Dual Basis]
Let $V$ be a finite dimensional vector space over $\F$ with a basis $B = \{v_1, \dots, v_n\}$.
Then we define the \vocab{dual basis} for $V^*$ to be $B^* = \{\varepsilon_1, \dots, \varepsilon_n\}$ where $\varepsilon_j(v_i) = \delta_{ij}$.
In particular, if $v \in V$ can be written as $v = \sum_{i = 1}^n \lambda_i v_i$, then $\varepsilon_j(v) = \lambda_j$.
\end{definition}
Since this is a basis of $V^*$, we can immediately see that $\dim V^* = \dim V$. Sometimes it's also convenient to think of $V^*$ as the space of row vectors of length $m$ over $\F$\footnote{There's also a hidden inner product structure here, but we won't dwell on that.}
\subsection{The Annihilator}
When we have some subset $U$ of a vector space $V$, we frequently care about the set of linear maps that are zero on everything in $U$. This construction ends up gives us a subspace of the dual space $V^*$.
\begin{definition}[Annihilator]
If $V$ is a vector space over $\F$, and $U \subseteq V$, then the \vocab{annihilator} of $U$ is the set $U^0$ of linear maps $V \rightarrow \F$ that are zero on $U$,
$$
U^0 = \{ \alpha \in V^* \mid \alpha(u) = 0 \text{ for all }u \in U\}.
$$
\end{definition}
\begin{lemma}[Annihilator is a Subspace]
The annihilator is always a subspace of the dual, that is, $U^0 \leq V^*$.
\end{lemma}
\begin{proof}[Proof Sketch]
Check definitions.
\end{proof}
By thinking about the annihilator, we also get a result that looks a little like rank-nullity.
\begin{lemma}[Rank-Nullity for the Annihilator]
If $V$ is a finite dimensional vector space and $U \leq V$, then $\dim V = \dim U + \dim U^0$.
\end{lemma}
\begin{proof}[Proof Sketch]
Let $\{v_1, \dots, v_k\}$ be a basis of $U$, and extend it to a basis $B = \{v_1, \dots, v_n\}$ of $V$. Then the dual basis is $B' = \{\varepsilon_1, \dots, \varepsilon_n\}$.
We then just need to check that $\{\varepsilon_{k + 1}, \dots, \varepsilon_n\}$ is a basis of $U^0$.
\end{proof}
\subsection{The Dual Map}
We will now see that thinking about the dual will give us a basis-free interpretation of what corresponds with the transpose of a matrix.
\begin{definition}[Dual Map]
Let $V$ and $W$ be vector spaces, and suppose that $\alpha: V \rightarrow W$ is a linear map. Then we define the \vocab{dual map} of $\alpha$ to be linear map $\alpha^*: W^* \rightarrow V*$ given by $\varepsilon \mapsto \varepsilon \circ \alpha$.
\end{definition}
To see what this is all about, consider the following example.
\begin{example}[Using the Dual Map]
Let $V$ and $W$ be $\R$ vector spaces, and let $B = \{v_1, v_2, v_3\}$ be a basis for $V$ and $C = \{w_1, w_2\}$ be a basis for $W$.
We define the linear map $\alpha: V \rightarrow W$ by
\begin{align*}
\alpha(v_1) &= w_1 + w_2 \\
\alpha(v_2) &= 2w_1 - w_2 \\
\alpha(v_3) &= 6w_1 + 3w_2.
\end{align*}
Now we can see what the dual map $\alpha^*: W^* \rightarrow V^*$ looks like. Let $B^* = \{v_1^*, v_2^*, v_3^*\}$ be the corresponding dual basis of $V$, and let $C^* = \{w_1^*, w_2^*\}$ be the corresponding dual basis of $W$. Then for any $v = av_1 + bv_2 + cv_3$ we have
\begin{align*}
[\alpha^*(w_1^*)](v) &= w_1^*(\alpha(av_1 + bv_2 + cv_3)) \\
&= w_1^*((a + 2b + 6c)w_1 + (a - b + 3c)w_2)\\
&= a + 2b + 6c,
\end{align*}
and in this way, we can really write
$$
\alpha^*(w_1^*) = v_1^* + 2v_2^* + 6v_3^*, \quad \text{and} \quad \alpha^*(w_2^*) = v_1^* - v_2^* + 3v_3^*,
$$
which totally specifies $\alpha^*$.
\end{example}
A slightly more interesting way of viewing this example is through the lense of matrices.
\begin{example}[Above with Matrices]
Continuing on from the example above, we can see that in the $B, C$ basis we have
$$
[\alpha]_{B, C} = \begin{pmatrix}
1 & 2 & 6 \\
1 & -1 & 3
\end{pmatrix}.