-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathmain.tex
896 lines (781 loc) · 54.1 KB
/
main.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
\documentclass[twoside, titlepage]{article}
\input{preamble.tex}
\title{%
Summary \\
\large WI4425 - Financial Markets Theory
}
\author{Auke Schaap}
\date{June 2023}
\begin{document}
\maketitle
\newpage
\thispagestyle{empty}
\begin{table}
\centering
\begin{tabular}{l|llll}
Lecture & Reading & Exercises & Document \\
\hline
1 & 1.1, 1.2 & 1 \& 2 & FMT 1 \\
2 & 1.3, 2.1 & 3 \& 5 & FMT 1 \\
3 & 2.1, 2.2 & 4, 1 & FMT 1, FMT 2 \\
4 & 2.2, 2.3 & 2 & FMT 2 \\
5 & 2.3 & 3.6 \& 3.8 & Book \\
6 & 2.3, 2.4 & 1 & Insurance Dominance \\
7 & 3.1, 3.2 & 2 & Insurance Dominance \\
8 & 3.3, 3.4, 5.1 & All & Portfolio \\
9 & 5.2, 5.3 & All & Risk neutral measure \\
10& 4.4 & All & Risk neutral measure
\end{tabular}
\caption{Overview of the lectures, readings, exercises and documents.}
\end{table}
\begin{table}
\centering
\begin{tabular}{l|l}
Chapter & Sections \\
\hline
1 & 1.1, 1.2, 1.3 \\
2 & 2.1, 2.2, 2.3, 2.4 \\
3 & 3.1, 3.2, 3.3, 3.4 \\
4 & 4.4 \\
5 & 5.1, 5.2, 5.3
\end{tabular}
\caption{Overview of the chapters and sections.}
\end{table}
\clearpage
\thispagestyle{empty}
\vspace*{\fill}
\section*{Disclaimer}
This document is not written by the lecturer, but by a student. It is not guaranteed to be free of errors, nor is it guaranteed to be complete. It is not a replacement for the lectures, nor for the book. It is merely a summary of the lectures and the book, and is intended to be used as a study aid. Use at your own risk.
\bigskip
\begin{flushright}
\textit{Auke Schaap, \today}
\end{flushright}
\vspace*{\fill}
\newpage
\thispagestyle{empty}
\tableofcontents
\clearpage
\lecture{1}
\section{Choices under certainty}
The couple $(X, \mathcal{R})$, represents the \textit{agent's choice problem}, where
\begin{itemize}
\item $X$ is a set of choices, which, in the classical consumption problem, concerns bundles of goods. $X$ is therefore a subset of $\R^L$, where $L \in \N$ is the number of goods.
\item $\mathcal{R}$ is a (weak) \textit{preference relation} on $X$. Given $x,y \in X$, if $x \mathcal{R} y$ we say that the bundle of goods $x$ is at least as preferred as the bundle of goods $y$.
\end{itemize}
From this we can derive a two other preference relations. The \textit{strong preference relation} $\mathcal{P}$ is defined as $x \mathcal{P} y \iff x \mathcal{R} y$ \text{but not} $y \mathcal{R} x$. The \textit{indifference relation} $\mathcal{I}$ is defined as $x \mathcal{I} y \iff x \mathcal{R} y $ and $y \mathcal{R} x$.
\begin{assumption}[Rationality] \label{ass:rationality}
The preference relation $\mathcal{R}$ is said to be \textit{rational} if it satisfies:
\begin{itemize}
\item \textit{Reflexivity}: $x \mathcal{R} x$ for all $x \in X$.
\item \textit{Completeness}: For all $x, y \in X$, either $x \mathcal{R} y$ or $y \mathcal{R} x$.
\begin{explanation}
This means that the agent is able to compare any two bundles of goods.
\end{explanation}
\item \textit{Transitivity}: For all $x, y, z \in X$, if $x \mathcal{R} y$ and $y \mathcal{R} z$, then $x \mathcal{R} z$.
\begin{explanation}
This means that the agent is consistent.
\end{explanation}
\end{itemize}
\end{assumption}
\begin{assumption}[Continuity] \label{ass:continuity}
The preference relation $\mathcal{R}$ is said to be \textit{continuous} if for all $x \in X$, the sets $\{y \in X : y \mathcal{R} x\}$ and $\{y \in X : x \mathcal{R} y\}$ are closed.
\end{assumption}
Introducing a utility function $u : X \to \R$ that represents the preference relation $\mathcal{R}$ if
\[
x \mathcal{R} y \iff u(x) \geq u(y) \; \text{ for all } x, y \in X,
\]
we can write the agent's choice problem as the couple $(X, u)$.
\begin{theorem}
Let $\mathcal{R}$ be a rational and continuous preference relation on $X$. Then there exists a continuous utility function $u : X \to \R$ that represents $\mathcal{R}$.
\end{theorem}
Note that this does not say anything about the uniqueness of the utility function. If $u$ is a utility function that represents $\mathcal{R}$, then so does any strictly increasing transformation of $u$.
\begin{assumption}
We furthermore assume that:
\begin{itemize}
\item $\mathcal{R}$ is strictly monotone if, for every $x, y \in X$ such that $y > x$, $x \mathcal{P} y$.
\item $\mathcal{R}$ is strictly convex if, for every $x, y, z \in X$ with $x \not = y$ such that $x \mathcal{R} z$ and $y \mathcal{R} z$, for all $\alpha \in (0, 1)$, $\alpha x + (1 - \alpha) y \mathcal{P} z$.
\end{itemize}
\end{assumption}
If we assume the utility functions to be twice continuously differentiable, then the first partial derivative of a utility function, the \textit{marginal utility}, is positive, and the second partial derivative is negative.
We assume that agents are characterized by \textit{substantial rationality}: An agent pursues his goals in the most appropriate way, under the constraints imposed by the environment. We, furthermore, dip the agent in a perfectly competitive market. This means that the agent is a \textit{price taker}. Given the wealth $w \in \R_{\geq0}$ and the price vector $p \in \R_{\geq 0}^L$, the set on which an agent then makes its choices (the \textit{budget constraint}) is
\[
B := \left\{x \in X : p^\top x \leq w\right\},
\]
for some closed set $X \subset \R_{\geq0}^L$. Given the couple $(X, u)$ and the budget constraint $B$, the \textit{optimal choice} is
\[
x^* := \max_{x \in B} u(x).
\]
If the utility function is continuous, then there always exists a solution to this optimization problem, if $p \in \R_{>0}^L$.
\begin{explanation}
As the utility function is considered to be strictly monotone, it is always optimal for an agent to spend all his wealth. Thus, the budget constraint becomes an equality constraint.
\end{explanation}
We can solve it by the Lagrangian method. The Lagrangian is:
\[
\mathcal{L}(x, \lambda) = u(x) + \lambda (w - p^\top x).
\]
This means that we solve the set of equations:
\begin{align*}
\frac{\partial u}{\partial x_l}\left(x^*\right) &= \lambda p_l, \; \forall l = 1, \dots, L, \\
w &= p^\top x^*.
\end{align*}
We define the \textit{indirect utility function} as $v(\omega) \equiv u(x^*)$.
Suppose we endow agent $i$ with endowment vector $e^i \in \R_{\geq0}^L$, and let $x(p)$ be the optimal consumption vector for price $p$. We can then write the demand $z^i(p)$ as
\[
z^i(p) = x(p) - e^i.
\]
\begin{definition}[Internal consistency requirement]
We require that agents optimize their utility functions under the budget constraint.
\end{definition}
\section{General Equilibrium Theory}
We have an economy with $I>1$ agents and $L>1$ goods. Every agent has solved his optimal consumption problem for his given wealth, $\omega^i = p^\top e^i$. The demand function summarizes his choices.
\begin{definition}[External consistency requirement]
We solve the market by putting demand equal to supply:
\[
z(p) \equiv \sum^I_{i=1}(x^i(p) - e^i) = \mathbf{0}
\]
\end{definition}
The price vector that solve this equation is the \textit{equilibrium price vector}. The couple $(p^*, x^*)$ is the \textit{equilibrium allocation}.
\begin{theorem}
Let the preference relation $\mathcal{R}^i$ be rational, continuous, strictly monotone and strictly convex for all $i = 1, \dots, I$, and suppose that $\sum^I_{i=1}e^i \in \R_{>0}^L$. Then there exists a unique equilibrium price vector $p^* \in \R_{>0}^L$.
\end{theorem}
\lecture{2}
\section{Pareto optimality}
Given an economy described by $I$ couples $(\mathcal{R}^i, e^i)$, we define an allocation $x$ by:
\[
x = (x^1, \dots, x^I) \;\; x^i \in \R_{\geq0}^L.
\]
This allocation is \textit{feasible} if we have
\[
\sum^I_{i=1}x^i \leq \sum^I_{i=1}e^i.
\]
Given two feasible allocations $x$ and $y$, then $x$ \textit{Pareto dominates} $y$ if and only if
\begin{align*}
x^i \mathcal{R} y^i \forall i = 1, \dots, I, \\
\exists\; 1 \leq j \leq I : x^i \mathcal{P} y^i.
\end{align*}
\begin{explanation}
$x$ is at least as preferred as $y$ for all agents, and strictly preferred by at least one agent. Thus, it is a better alternative.
\end{explanation}
A feasible allocation $x$ is an \textit{efficient allocation}, or \textit{Pareto optimal allocation}, if there is no other feasible allocation $y$ such that $y$ Pareto dominates $x$.
\begin{theorem}[First Welfare Theorem]
If $(p^*, x^*)$ is a competitive equilibrium (and the preference relations of the agents are strictly monotone), then the allocation $x^*$ is Pareto optimal.
\begin{explanation}
So, there is an invisible hand in the market: agents maximizing their own utility in a perfectly competitive market lead to a socially optimal result.
\end{explanation}
\end{theorem}
\begin{theorem}[Second Welfare Theorem]
Let $x^*$ be a Pareto optimal allocation, $x^{i*} \in \R_{\geq0}^L$. If the preference relations $\mathcal{R}^i$ are rational, continuous, strictly monotone and strictly convex, then $x^*$ is the competitive equilibrium allocation, given that the initial endowments are $e^i = x^{i*}$.
\begin{explanation}
We can reach any Pareto optimal allocation by redistributing the goods in a proper way.
\end{explanation}
\end{theorem}
Consider an economy $(u^i, e^i)$ and an associated competitive equilibrium $(p^*, x^*)$. Now consider an agent who has an optimal choice given the equilibrium price vector $p^*$ equal to the initial aggregate endowment. The agent has utility function
\begin{align*}
\textbf{u}(x) := \max_{x \in \R_{\geq0}^L, i=1, \dots, I} \sum^I_{i=1}a^i u^i(x^i) \;\; \\ \text{s.t.} \sum^I_{i=1}x^i_l \leq \sum^I_{i=1}x_l.
\end{align*}
This is the \textit{social welfare function}. If the equilibrium price vector does not depend on the weights, we say that the economy satisfies the \textit{aggregation property}. If an economy satisfies the aggregation propery, then the economy with $I$ agents and the economy with the one representative agent are observationally equivalent.
\section{Choices Under Risk}
We now consider a state of the world at time $t=0$ and at $t=1$. The beliefs are modeled by a probability space $(\Omega, \mathcal{F}, \mu)$. We distinguish between:
\begin{itemize}
\item \textit{Risk}: The probabilities are known.
\item \textit{Uncertainty}: The probabilities are unknown.
\end{itemize}
Acts are modeled by random variables $\tilde{x}: \Omega \rightarrow \R$, where we assume $\tilde{x}$ is measurable in the following sense:
\[
\forall B \in \mathcal{B}(\R) : \;\; \{ \omega \in \Omega : \tilde{x}(\omega) \in B \} \in \mathcal{F}.
\]
$\mu$ induces a probability measure $\pi$ on $\R$, $\pi(B) = \mu(\tilde{x}^{-1}(B))$, with $B \in \mathcal{B}(\R)$. The acts attain a finite set of values, specified a priori: $ [ x_1, \dots, x_S ], \; x_i < x_{i+1} \in \R, \;\; i = 1, \dots, S - 1$. We now have a set of acts, or \textit{gambles} $\mathcal{M}$, which are r.v.'s having support on the $x_i$ identified by a probability distribution:
\[
\tilde{x} \in \mathcal{M} \iff \tilde{x} \equiv [x_1, \dots, x_S ; \pi_1, \dots, \pi_S] \;\; \text{s.t.} \;\; \pi_s \geq 0, \sum^S_{s=1}\pi_s = 1.
\]
\lecture{3}
\section{Expected Utility}
With the the characterization $(\mathcal{M}, \mathcal{R})$, we introduce a function of $\tilde{x}$ that represents $\mathcal{R}$, the \textit{expected utility}. We again assume rationality and continuity, but also independence:
\begin{assumption}[Independence]
The preference relation $\mathcal{R}$ satisfies the \textit{independence} assumption, if for all $\tilde{x}_1, \tilde{x}_2, \tilde{x}_3 \in \mathcal{M}$ and $\alpha \in [0,1]$, we have
\[
\tilde{x}_1 \mathcal{R} \tilde{x}_2 \iff \alpha \tilde{x}_1 + (1 - \alpha)\tilde{x}_3 \mathcal{R} \alpha \tilde{x}_2 + (1 - \alpha)\tilde{x}_3.
\]
\end{assumption}
We can now define the \textit{expected utility} by:
\[
U(\tilde{x}) := \sum^S_{s=1}\pi_s u(x_s) = \E[u(\tilde{x})],
\]
where $u$ is the utility function that represents the preference relation with certainty.
\begin{theorem}[Theorem 2.1] % Theorem 2.1
If $\mathcal{R}$ satisfies the assumptions, then there exist $S$ scalars $u(x_s)$ such that:
\[
\forall \tilde{x}_1, \tilde{x}_2 \in \mathcal{M} : \;\; \tilde{x}_1 \mathcal{R} \tilde{x}_2 \iff U(\tilde{x}_1) \geq U(\tilde{x}_2).
\]
\end{theorem}
\begin{proof}
see Barucci, Fontana, 2017, p. 21.
\end{proof}
\section{Risk Aversion}
A gamble $\tilde{x}$ is an \textit{actuarially fair} gamble if:
\[
\sum^S_{s=1}\pi_s x_s = 0.
\]
We have the following types of agents:
\begin{itemize}
\item \textit{Risk averse}: Agent does not accept or is indifferent to any actuarially fair gamble at all wealth levels.
\item \textit{Risk neutral}: Agent is indifferent to any actuarially fair gamble.
\item \textit{Risk loving}: Agent does accept any actuarially fair gamble.
\end{itemize}
The \textit{risk premium} of a gamble $\tilde{x}$ is the amount an agent is willing to pay to exchange his expected wealth for certain wealth, i.e. to avoid the gamble. That is, the amount $\rho_u(\tilde{x})$ such that:
\begin{align*}
u\left(\E[\tilde{x}] - \rho_u(\tilde{x})\right) = \E[u(\tilde{x})] \\[10pt]
\rho_u(\tilde{x}) := \max \{x \in \R_{\geq0} : u(\E[\tilde{x}] - x) = U(\tilde{x})\}
\end{align*}
The \textit{certainty equivalent} of a gamble $\tilde{x}$ is the amount of money an agent wants, to exchange his investment:
\[
\text{CE}_u(\tilde{x}) := \E[\tilde{x}] - \rho_u(\tilde{x}).
\]
\begin{proposition}
The following are equivalent:
\begin{enumerate}[(i)]
\item The agent is risk-averse.
\item $u$ is concave.
\item $\text{CE}_u(\tilde{x}) \leq \E[\tilde{x}]$.
\item $\rho_u(\tilde{x}) \geq 0$.
\end{enumerate}
\end{proposition}
Let $\tilde{x} = x + \tilde{\epsilon}$, where $\tilde{\epsilon}$ is a random variable with mean zero and variance $\sigma^2$, so that $\E[\tilde{x}] = x$. Now we can write:
\[
u\left(\E[\tilde{x}] - \rho_u(\tilde{x})\right) = u(x - \rho_u(\tilde{x})) = U(\tilde{x}).
\]
Using a Taylor expansion we find that $U(\tilde{x}) = \E[u(x + \tilde{\epsilon})] \approx u(x) + \frac{\sigma^2}{2}u''(x)$, and that $u(x - \rho_u(\tilde{x})) \approx u(x) - u'(x)\rho_u(\tilde{x})$. Hence, $\rho_u(\tilde{x}) \approx \frac{1}{2} r^a_u(x)\sigma^2$, where
\[
r^a_u(x) := -\frac{u''(x)}{u'(x)},
\]
the coefficient of \textit{absolute risk aversion}. If we analyse the multiplicative noise $\tilde{x} = x(1 + \tilde{\epsilon})$, we find the coefficient of \textit{relative risk aversion}:
\[
r^r_u(x) := -\frac{u''(x)}{u'(x)}x.
\]
We use the following classification of risk-aversion:
\begin{itemize}
\item \textit{DARA}: Decreasing absolute risk aversion, $x \mapsto r^a_u(x)$ is a decreasing function.
\item \textit{CARA}: Constant absolute risk aversion, $x \mapsto r^a_u(x)$ is a constant function.
\item \textit{IARA}: Increasing absolute risk aversion, $x \mapsto r^a_u(x)$ is an increasing function.
\end{itemize}
We say of two agents $a$ and $b$ that $a$ is more risk averse than $b$ if agent $b$ always accepts a gamble if agent $a$ does, or if for every gamble, the risk premium of agent $a$ is greater than or equal to the risk premium of agent $b$.
\lecture{4}
\begin{proposition}[Proposition 2.2] % Proposition 2.2
Given two increasing and strictly concave utility functions $u^a$ and $u^b$, the following conditions are equivalent:
\begin{itemize}
\item $r^a_{u^a}(x) \geq r^a_{u^b}(x)$ for all $x \in \R_{\geq0}$.
\item There exists an increasing, concave function $g$ such that $u^a(x) = g(u^b(x))$ for all $x \in \R_{\geq0}$.
\item $u^a$ is more risk averse than $u^b$, i.e. $\rho_{u^a}(x + \tilde{\epsilon}) \geq \rho_{u^b}(x +\tilde{\epsilon})$ for all $x \in \R_{\geq0}$ and for any random variable $\tilde{\epsilon}$ with mean zero.
\end{itemize}
\end{proposition}
\begin{proof}
$(i) \iff (ii)$: Since $u^a$ and $u^b$ are both increasing, concave and twice differentiable, there exists a twice differentiable increasing function $g$ such that $u^a(x) = g(u^b(x))$ for all $x \in \R_{\geq0}$. Let $g := u^a \circ (u^b)^{-1}$, which is increasing and twice differentiable. Differentiating $u^a$ yields:
\begin{align*}
u^{a'} &= g'(u^b(x))u^{b'}(x) \\
u^{a''} &= g'(u^b(x))u^{b''}(x) + g''(u^b(x))(u^{b'}(x))^2
\end{align*}
Dividing by $u^{a'}$ yields:
\begin{align*}
\frac{u^{a''}}{u^{a'}} &= \frac{u^{b''}(x)}{u^{b'}(x)} + \frac{g''(u^b(x))}{g'(u^b(x))}(u^{b'}(x)) \\
-r^a_{u^a}(x) &= -r^a_{u^b}(x) + \frac{g''(u^b(x))}{g'(u^b(x))}(u^{b'}(x)) \\
r^a_{u^a}(x) &= r^a_{u^b}(x) - \frac{g''(u^b(x))}{g'(u^b(x))}(u^{b'}(x))
\end{align*}
Thus, with $u^b$ and $g$ increasing, $r^a_{u^a}(x) = r^a_{u^b}(x) \iff g''(u^b(x)) \leq 0$, for all $x \in \R_{\geq0}$.
$(ii) \implies (iii)$: Suppose that agents $a$ and $b$ are endowed with the same wealth $x \in \R$. For an arbitrary random variable $\tilde{\epsilon}$ with mean zero, we have that:
\begin{align*}
u^a(x -\rho_{u^a}(x + \tilde{\epsilon})) &= \E[u^a(x + \tilde{\epsilon})] \\
&= \E[g(u^b(x + \tilde{\epsilon}))] \\
&\leq g(U^b(x + \tilde{\epsilon})) \\
&= g(u^(x - \rho_{u^b}(x + \tilde{\epsilon}))) \\
&= u^a(x - \rho_{u^b}(x + \tilde{\epsilon})),
\end{align*}
where the inequality is due to Jensen's inequality, since $g$ is concave. Because $u^a$ is increasing, we then have $\rho_{u^a}(x + \tilde{\epsilon}) \geq \rho_{u^b}(x + \tilde{\epsilon})$.
$(iii) \implies (ii)$: As in the first part, since $u^a$ and $u^b$ are strictly increasing, ther exists an increasing function $g$ such that $u^a(x) = g(u^b(x))$. We will need to show that $g$ is concave. Let $x, x' \in \R$ and $\lambda \in [0, 1]$. Define $ \bar{x} := \lambda x + (1 - \lambda)x'$, and consider the gamble $\tilde{\epsilon}$ defined as follows:
\[
\tilde{\epsilon} = \begin{cases}
(1- \lambda)(x - x') & \text{with probability } \lambda \\
\lambda(x - x') & \text{with probability } 1 - \lambda
\end{cases}
\]
We have $\E[\tilde{\epsilon}] = 0$. Then, by the definition of risk premium
\begin{align*}
\lambda u^b(x) + (1 - \lambda)u^b(x') &= \E[u^b(\bar{x} + \tilde{\epsilon})] \\
&= u^b(\bar{x} - \rho_{u^b}(\bar{x} + \tilde{\epsilon}))
\end{align*}
Since $u^a(x) = g(u^b(x))$,
\[
g(\lambda u^b(x) + (1 - \lambda)u^b(x')) = u^a(\bar{x} - \rho_{u^b}(\bar{x} + \tilde{\epsilon}))
\]
On the other hand,
\begin{align*}
\lambda g(u^b(x)) + (1 - \lambda)g(u^b(x')) &= \lambda u^a(x) + (1 - \lambda)u^a(x') \\
&= \E[u^a(\bar{x} + \tilde{\epsilon})] \\
&= u^a(\bar{x} - \rho_{u^a}(\bar{x} + \tilde{\epsilon}))
\end{align*}
Since $u^a$ is increasing, and using property $(iii)$, we get
\[
g(\lambda u^b(x) + (1 - \lambda)u^b(x')) \geq \lambda g(u^b(x)) + (1 - \lambda)g(u^b(x'))
\]
As $x, x' \in \R$ and $\lambda \in [0, 1]$ were arbitrary, $g$ is concave.
\end{proof}
Examples of utility functions are:
\begin{itemize}
\item Exponential utility functions: $u(x) = -e^{-\alpha x}$, where $\alpha > 0$.
\begin{explanation}
The coefficient of absolute risk aversion is $r^a_u(x) = \alpha$, which is constant (and positive). Therefore, the exponential utility function is CARA and the coefficient of relative risk aversion $r^r_u(x) = \alpha x$ is increasing.
\end{explanation}
\item Quadratic utility functions: $u(x) = x - \frac{b}{2}x^2$.
\begin{explanation}
The coefficient of absolute risk aversion is $r^a_u(x) = \frac{b}{1 - bx}$, which is increasing and positive. Therefore, the quadratic utility function is IARA and the coefficient of relative risk aversion $r^r_u(x) = \frac{bx}{1 - bx}$ is increasing.
\end{explanation}
\item Power utility functions: $u(x) = \frac{b}{b-1}x^{1 - \frac{1}{b}}$, where $b > 0$.
\begin{explanation}
The coefficient of absolute risk aversion is $r^a_u(x) = \frac{1}{bx}$, which is decreasing and positive. Therefore, the power utility function is DARA and the coefficient of relative risk aversion $r^r_u(x) = \frac{1}{b}$ is constant.
\end{explanation}
\item Logarithmic utility functions: $u(x) = \ln(bx)$, where $b > 0$.
\begin{explanation}
The coefficient of absolute risk aversion is $r^a_u(x) = \frac{1}{x}$, which is decreasing and positive. Therefore, the logarithmic utility function is DARA and the coefficient of relative risk aversion $r^r_u(x) = 1$ is constant.
\end{explanation}
\end{itemize}
\section{Portfolio Problem}
An agent with initial wealth $W_0$ faces the following problem. Allocate $W_0$ among $N+1$ assets, of which $N$ are risky. The return of asset $n$ at time 1 is described by $\tilde{r}_n$, with $n = 1, \dots, N$. Asset 0 is the risk-free asset with return $r_f$. There are $S$ possible outcomes:
\[
\begin{pmatrix}
r_{11} & r_{12} & \dots & r_{1N} \\
r_{21} & r_{22} & \dots & r_{2N} \\
\vdots & \vdots & \ddots & \vdots \\
r_{S1} & r_{S2} & \dots & r_{SN}
\end{pmatrix}
\]
The agent invests the vector with amounts $\mathbf{w} = (w_1, \dots, w_N)$, $\mathbf{w} \in \R^{N}$. The budget constraint forces the agent to invest the rest in the risk-free asset:
\[
W_0 - \sum_{n=0}^N w_n.
\]
The wealth at time 1 becomes $y = R\mathbf{w}^\top$. The \textit{consumption plans} $\mathbf{c}$ reachable in $\R^S$ are given by the image of the transformation $R$, written as $I(R)$. A market is complete if every consumption plan is reachable, i.e. $I(R) = \R^S$. The $N$ assets are said to be \textit{non redundant} if their returns are linearly independent, i.e. $N \leq S$ and $\text{rank}(R) = N$. The \textit{uniqueness of representation} property states that for all $\mathbf{c} \in I(R)$ there exists a unique $\mathbf{w} \in \R^N$ such that $\mathbf{c} = R\mathbf{w}^\top$.
\lecture{5}
Assuming $R$ is full rank, the following cases are possible:
\begin{itemize}
\item $N > S$: $I(R) = \R^S$, and $\dim\{\mathbf{w} : R\mathbf{w}^\top = \mathbf{c}\} = N-S$.
\begin{explanation}
The market is complete, as every possible consumption plan is reachable ($I(R) = \R^S$). Some assets are \textit{redundant}. The number of redundant assets is $N-S$, and the number of non-redundant assets is $S$.
\end{explanation}
\item $N = S$: $I(R) = \R^S$, and $\dim\{\mathbf{w} : R\mathbf{w}^\top = \mathbf{c}\} = 0$.
\begin{explanation}
The market is complete, as every possible consumption plan is reachable ($I(R) = \R^S$). There are no redundant assets, as the number of non-redundant assets is $S$.
\end{explanation}
\item $N < S$: $I(R) \subset \R^S$, and $\dim\{\mathbf{w} : R\mathbf{w}^\top = \mathbf{c}\} = 0$.
\begin{explanation}
The market is incomplete, as not every possible consumption plan is reachable ($I(R) \subset \R^S$). There are no redundant assets, as the number of non-redundant assets is $S$.
\end{explanation}
\end{itemize}
The wealth at $t_1$ associated with the portfolio is given by:
\begin{align*}
\tilde{W} &= \left(W_0 - \sum_{n=1}^N w_n\right)r_f + \sum_{n=1}^N w_n \tilde{r}_n \\
&= W_0r_f + \sum_{n=1}^N w_n(\tilde{r}_n - r_f).
\end{align*}
An agent would maximize his expected utility given an initial wealth. Hence, the agent's \textit{optimal portfolio problem} is given by:
\[
\max_{\mathbf{w}} \E[u(\tilde{W})], \;\;\; \mathbf{w} \in \R^N.
\]
The first order necessary condition is given by $\E[u'(\tilde{W})(\tilde{r}_n - r_f)] = 0$, with $n = 1, \dots, N$. If $u'' < 0$, then there is a \textit{unique wealth value} $W$ that yield the maximum of the utility functions. This will be attained by a unique portfolio if the assets statisfy the uniqueness of representation property. The \textit{return risk premium} is defined as:
\[
\E[\tilde{r}_n - r_f].
\]
\begin{proposition}
Consider a strictly risk averse agent with a strictly increasing utility function. A sufficient condition such that $\mathbf{w}^* \not\in \R^N_{\geq 0}$ is that the return risk premium of all assets is non positive, and negative for at least one asset.
\begin{explanation}
As the agent is risk averse, he requires a positive risk premium to invest in a risky asset. If the risk premium is zero, the agent will not invest in the asset. If the risk premium is negative, the agent will short the asset. Hence, $\mathbf{w}^* \in \R^N_{\leq 0}$.
\end{explanation}
\end{proposition}
\begin{proposition}
A necessary and sufficient conditions so that $\mathbf{w}^* = 0$ for an agent with a strictly increasing and strictly concave utility function is that
\[
\E[\tilde{r}_n] = r_f, \;\;\; n = 1, \dots, N.
\]
\begin{explanation}
The agent is risk averse, due to the concavity of the utility function. Hence, the agent will only invest in a risky asset for a positive risk premium. If the risk premium is zero, the agent will invest in the risk-free asset, so $\mathbf{w}^* = 0$.
\end{explanation}
\end{proposition}
\begin{proposition}[Proposition 2.5] % Proposition 2.5
Given $N$ risky assets with returns $\tilde{r}_1, \dots, \tilde{r}_N$, and a risk-free asset with return $r_f > 0$. Let $u$ be differentiable, strictly increasing and concave. Then for $n = 1, \dots, N$:
\begin{align*}
\E[\tilde{r}_n - r_f] &\geq 0 \iff cov\left(\tilde{r}_n, u'(W_0\tilde{r}^*)\right) \leq 0, \\
\E[\tilde{r}_n - r_f] &\leq 0 \iff cov\left(\tilde{r}_n, u'(W_0\tilde{r}^*)\right) \geq 0.
\end{align*}
\begin{explanation}
A risk-averse agent wants to reduce the variance of his portfolio. For assets that do this (i.e. negative correlation between asset returns and wealth), a negative risk premium is accepted. For assets that increase the variance of the portfolio, a positive risk premium is required.
\end{explanation}
\end{proposition}
\begin{proof}
By the first order necessary condition, $\E[u'(\tilde{W})(\tilde{r}_n - r_f)] = 0$ for $n = 1, \dots, N$ and recalling that for two random variables $\tilde{x}_1, \tilde{x}_2$ it holds that $cov(\tilde{x}_1, \tilde{x}_2) = \E[\tilde{x}_1\tilde{x}_2] - \E[\tilde{x}_1]\E[\tilde{x}_2]$, we have:
\begin{align*}
cov(u'(\tilde{W}^*), \tilde{r}_n) &= \E[u'(\tilde{W}^*)\tilde{r}_ n] - \E[\tilde{r}_n]\E[u'(\tilde{W}^*)] \\
cov(u'(\tilde{W}^*), \tilde{r}_n) + \E[\tilde{r}_n]\E[u'(\tilde{W}^*)] &= \E[u'(\tilde{W}^*)\tilde{r}_n] \\
cov(u'(\tilde{W}^*), \tilde{r}_n) + \E[\tilde{r}_n]\E[u'(\tilde{W}^*)] &= r_f\E[u'(\tilde{W}^*)] \\
\frac{cov(u'(\tilde{W}^*), \tilde{r}_n)}{\E[u'(\tilde{W}^*)]} + \E[\tilde{r}_n] &= r_f \\[10pt]
\E[\tilde{r}_n] - r_f &= - \frac{cov(u'(\tilde{W}^*), \tilde{r}_n)}{\E[u'(\tilde{W}^*)]}, \;\;\; \text{for all } n = 1, \dots, N.
\end{align*}
As $u$ is increasing, $u'(\tilde{W}^*) > 0$ and hence the sign of the covariance determines the sign of the risk premium.
\end{proof}
For returns with a small dispersion, the following approximation holds:
\[
\E[\tilde{r}_n - r_f] \approx - \frac{u''\left(\E[\tilde{W}^*]\right)}{u'\left(\E[\tilde{W}^*]\right)} cov\left(\tilde{r}_n, \tilde{W}^*\right)
\]
\lecture{6}
\begin{proposition}
If the returns $\tilde{r}_n$, with $n = 1, \dots, N$, are iid, then the optimal portfolio for a risk averse agent is given by:
\[
w_n = \frac{W_0}{N}.
\]
\end{proposition}
\begin{proposition}
If an agent is risk averse with a strictly increasing utility function, then
\[
\E[\tilde{r}_n] - r_f \geq 0.
\]
\end{proposition}
Limit our portfolio problem to one risky and one risk-free asset. By $w(x)$ we denote the risky asset demand of an agent with initial wealth $x$.
\begin{proposition}
Let $u$ be an increasing, three times differentiable and strictly concave utility function. Then:
\begin{align*}
r^{a'}_u(z) < 0 \;\;\; \forall z \in \R_{\geq 0} & \implies w'(x) > 0 \;\;\; \forall x \in \R_{\geq 0} \\
r^{a'}_u(z) > 0 \;\;\; \forall z \in \R_{\geq 0} & \implies w'(x) < 0 \;\;\; \forall x \in \R_{\geq 0} \\
r^{a'}_u(z) = 0 \;\;\; \forall z \in \R_{\geq 0} & \implies w'(x) = 0 \;\;\; \forall x \in \R_{\geq 0}
\end{align*}
\begin{explanation}
DARA agents will increasingly invest in risky assets. IARA agents will decreasingly invest in risky assets. CARA agents will invest a constant proportion of their wealth in risky assets.
\end{explanation}
\end{proposition}
The absolute risk aversion coefficient gives information of changes in the levels of the risky asset demand, not in the percentage of wealth invested in the risky asset. For that, we use the relative risk aversion coefficient.
\begin{proposition}
Let $u$ be an increasing, three times differentiable and strictly concave utility function. Then:
\begin{align*}
r^{r'}_u(z) < 0 \;\;\; \forall z \in \R_{\geq 0} & \implies \frac{dw}{dx}\frac{x}{w} < 1 \;\;\; \forall x \in \R_{\geq 0} \\
r^{r'}_u(z) > 0 \;\;\; \forall z \in \R_{\geq 0} & \implies \frac{dw}{dx}\frac{x}{w} > 1 \;\;\; \forall x \in \R_{\geq 0} \\
r^{r'}_u(z) = 0 \;\;\; \forall z \in \R_{\geq 0} & \implies \frac{dw}{dx}\frac{x}{w} = 1 \;\;\; \forall x \in \R_{\geq 0}
\end{align*}
\end{proposition}
\section{Insurance, Demand and Prudence}
An \textit{Arrow} security delivers at time $t=1$ exactly one unit of wealth if an elementary event is realized, and zero otherwise. Suppose a strictly risk averse agent endowed in $t=0$ with an amount of money $W_0>0$ faces a possible monetary loss $D>0$ at time $t=1$, where the loss occurs with probability $0 < \pi < 1$. The expected utility is given by:
\[
\pi u(W_0 - D) + (1-\pi)u(W_0).
\]
Suppose the agent can buy an Arrow security that pays one unit of wealth in case the loss occurs. The price of this unit of wealth is $q>0$. The agent must solve the following optimization problem:
\[
\max_{q} \pi u(W_0 - D - wq + w) + (1-\pi)u(W_0 - wq), \;\;\; w \in \R_{\geq 0}.
\]
Here $w^* \geq 0$ is a solution of the problem if and only if:
\[
\pi(1-q) u'(W_0 - D - w^*q + w^*) - q(1-\pi)u'(W_0 - w^*q) \leq 0,
\]
with equality for $w^* > 0$. The Arrow security is in fact a gamble: $[1-q, -q; \pi, 1-\pi]$. The gamble is actuarially fair for $q = \pi$. So, for $w^* > 0$ we have
\[
u'\left(W_0 - D + w^*(1-q)\right) = u'\left(W_0 - w^*q\right).
\]
By concavity of $u$ we find $w^* = D$, so for actuarially fair price, the risk averse agent will insure themself completely. If the gamble is not fair, then there is partial insurance in case $q > \pi$, and over insurance in case $q < \pi$.
Consider now the generic gamble $\tilde{x} = [x_1, x_2; \pi, 1-\pi]$. The ratio of the marginal utilities in two states of the world, weighted by the probability of that state, denoted $\text{SMS}(x_1, x_2)$, is called the \textit{marginal rate of substitution}. This is equal to minus the slope of the tangent line to the indifference curve at the point $(x_1, x_2)$:
\[
\text{SMS}(x_1, x_2) = \left.-\frac{dx_2}{dx_1}\right|_{U(x) = \tilde{U}} = - \frac{\pi u'(x_1)}{(1-\pi)u'(x_2)}.
\]
\begin{figure}[ht]
\centering
\incfig[0.65]{insurance}
\caption{Expected utility in the state space. }
\label{fig:insurance}
\end{figure}
This allows us to show that the indifference curves in the state space are convex. It suffices to show that
\[
\frac{d\text{SMS}(x_1, x_2)}{dx_1} \leq 0, \;\;\; \forall x_1, x_2 \in \R^2_{\geq 0}.
\]
Now suppose you can buy Arrow securities paying 1 in case you end up in $x_1$, or $x_2$, with prices $q_1$ and $q_2$. Then the optimization problem becomes
\begin{align*}
\max_{x_1, x_2} U(x_1, x_2) = \pi u(x_1) + (1-\pi)u(x_2) \\
\text{s.t. } \; q_1x_1 + q_2x_2 \leq q_1e1 + q_2e2.
\end{align*}
Then we find as necessary condition to be satisfied in the solution:
\[
\frac{\pi u'(x^*_1)}{(1-\pi)u'(x^*_2)} = \frac{q_1}{q_2}.
\]
In case the securities are fairly priced, the agent will reach an endowment with no risk. Otherwise,
\begin{align*}
\frac{q_1}{q_2} > \frac{\pi}{(1-\pi)} & \implies x^*_1 < x^*_2 \\
\frac{q_1}{q_2} < \frac{\pi}{(1-\pi)} & \implies x^*_1 > x^*_2.
\end{align*}
\lecture{7}
\section{Stochastic Dominance}
Take the portfolio problem setup with $N$ risky assets, having random returns $\tilde{r}_i$, with covariance matrix
\[
V = \begin{pmatrix}
\sigma^2_1 & cov(\tilde{r}_1, \tilde{r}_2) & \cdots & cov(\tilde{r}_1, \tilde{r}_N) \\
cov(\tilde{r}_2, \tilde{r}_1) & \sigma^2_2 & \cdots & cov(\tilde{r}_2, \tilde{r}_N) \\
\vdots & \vdots & \ddots & \vdots \\
cov(\tilde{r}_N, \tilde{r}_1) & cov(\tilde{r}_N, \tilde{r}_2) & \cdots & \sigma^2_N
\end{pmatrix}.
\]
Furthermore, we denote the relative weights by
\[
w \in \Delta_N, \;\;\; \Delta_N = \left\{x \in \R^N : \; \sum_{i=1}^N x_i = 1\right\}.
\]
Then, by denoting by $e$ the vector of expected returns of the random assets, we find for expected return and variance of the portfolio:
\begin{align*}
\E[\tilde{r}] = w^\top e = \sum^N_{i=1}w_ie_i = \sum^N_{i=1}w_i\E[\tilde{r}_i], \\
\sigma^2(w) = w^\top V w = \sum^N_{i=1}\sum^N_{j=1}w_icov(\tilde{r}_i, \tilde{r}_j)w_j.
\end{align*}
The matrix $V$ is symmetric. In case all returns are linearly independent, then the matrix $V$ is non singular and positive definite.
\begin{definition}[First order stochastic dominance]
A portfolio $w^1$ stochastically dominates $w^2$ according to the \textit{first order stochastic dominance} criterion ($w^1 \succeq_{FSD} w^2$) if and only if $U(\tilde{r}_1) \geq U(\tilde{r}_2)$ for all non-decreasing utility functions.
\end{definition}
For this proposition, we assume the random variables to be normalized to the interval $[0, 1]$, so $F_i(1) = 1$.
\begin{proposition}
The following statements are equivalent:
\begin{itemize}
\item $w^1 \succeq_{FSD} w^2$
\item $F_1(x) \leq F_2(x)$ for all $x \in [0, 1]$
\item $\tilde{r}^1 =^d \tilde{r}^2 + \tilde{\epsilon}$ where $\tilde{\epsilon}$ is a positive random variable.
\end{itemize}
\end{proposition}
This is relatively weak, as it is defined on a large class of utility functions.
\begin{definition}[Second order stochastic dominance]
A portfolio $w^1$ stochastically dominates $w^2$ according to the \textit{second order stochastic dominance} criterion ($w^1 \succeq_{SSD} w^2$) if and only if $U(\tilde{r}_1) \geq U(\tilde{r}_2)$ for all concave utility functions.
\end{definition}
\begin{proposition}
The following statements are equivalent:
\begin{itemize}
\item $w^1 \succeq_{SSD} w^2$
\item $\E[\tilde{r}^1] = \E[\tilde{r}^2]$, with $G(y) = \int_0^y\left(F_1(z) - F_2(z)\right) dz \leq 0$ for all $y \in [0, 1]$.
\item $\tilde{r}^2 =^d \tilde{r}^1 + \tilde{\epsilon}$, where $\tilde{\epsilon}$ is a random variable such that $\E[\tilde{\epsilon} | \tilde{r}^1] = 0$.
\end{itemize}
\end{proposition}
The result of this is that
\[
w^1 \succeq_{SSD} w^2 \implies \E[\tilde{r}^1] = \E[\tilde{r}^2] \text{ and } \sigma^2(w^1) \leq \sigma^2(w^2).
\]
We now have partial ordering, i.e. only for portfolios with the same expected return.
\begin{definition}[Second order stochastic monotonic dominance]
A portfolio $w^1$ stochastically dominates $w^2$ according to the \textit{second order stochastic monotonic dominance} criterion ($w^1 \succeq_{SSD}^M w^2$) if and only if $U(\tilde{r}_1) \geq U(\tilde{r}_2)$ for all non-decreasing and concave utility functions.
\end{definition}
\begin{proposition}
The following statements are equivalent:
\begin{itemize}
\item $w^1 \succeq_{SSD}^M w^2$
\item $\E[\tilde{r}^1] \geq \E[\tilde{r}^2]$, with $G(y) = \int_0^y\left(F_1(z) - F_2(z)\right) dz \leq 0$ for all $y \in [0, 1]$.
\item $\tilde{r}^2 =^d \tilde{r}^1 + \tilde{\epsilon}$, where $\tilde{\epsilon}$ is a random variable such that $\E[\tilde{\epsilon} | \tilde{r}^1] \leq 0$.
\item $\tilde{r}^2 =^d \tilde{r}^1 + \tilde{\xi} + \tilde{\nu}$, where $\tilde{\xi}$ is a positive random variable and $\E[\tilde{\nu} | \tilde{r}^1 + \tilde{\xi}] = 0$.
\end{itemize}
\end{proposition}
Note that, by definition, second order stochastic dominance implies second order stochastic monotonic dominance. That is, if $w^1 \succeq_{SSD} w^2$, then $w^1 \succeq_{SSD}^M w^2$.
\section{Mean-Variance Analysis}
The \textit{mean-variance} criterion is given by:
\[
w^1 \succeq_{MV} w^2 \iff \E[\tilde{r}^1] \geq \E[\tilde{r}^2] \text{ and } \sigma^2(w^1) \leq \sigma^2(w^2).
\]
\begin{explanation}
That is, an agent prefers random variables with higher expected values, and dislikes random variables with higher variance.
\end{explanation}
This is again a partial ordering; there is no ordering for portfolios with higher risk and higher expected return, or lower risk and lower expected return. The justification for the criterion is as follows. If $u$ is quadratic, then $w^1 \succeq_{MV} w^2 \iff U(w^1) \geq U(w^2)$. We can expand this up to the second order, and find that $\E[u(\tilde{r})] \approx u(\E[\tilde{r}]) + \frac{1}{2}u''(\E[\tilde{r}])\sigma^2(\tilde{r})$.
\lecture{8}
\section{Portfolio Frontier}
The \textit{porfolio frontier} (PF) is identified by the second order stochastic dominance criterion. Given a portfolio with expected return $\E[\tilde{r}^p]$. The portfolio $w^p$ that is not SSD by any other portfolios with that return, i.e. the portfolio with the smallest variance of all portfolios having that return, is part of the PF. We can find $w^p$ by solving:
\begin{align*}
\min_w& \;\frac{1}{2}w^\top V w, \;\;\; w \in \R^N \\
\text{s.t.}& \;\;\; w^\top e = \E[\tilde{r}^p], \;\;\; \sum_{i=1}^N w_i = 1.
\end{align*}
The solution to this is given by
\[
w^p = g + h \E[\tilde{r}^p],
\]
where $g$ and $h$ are not dependent on the expected return, and given by:
\begin{align*}
g = \frac{BV^{-1}\mathbf{1}-AV^{-1}e}{D}, && h = \frac{CV^{-1}e - AV^{-1}\mathbf{1}}{D},
\end{align*}
\vspace{-15pt}
\begin{align*}
A = \mathbf{1}^\top V^{-1}e, && B = e^\top V^{-1}e, && C = \mathbf{1}^\top V^{-1} \mathbf{1}, && D = BC - A^2.
\end{align*}
The PF has the following important properties:
\begin{enumerate}
\item Given two portfolios in the PF, then any other portfolio can be attained by a linear combination of those two portfolios.
\item The covariance and variance are given by:
\begin{align*}
cov(\tilde{r}^p, \tilde{r}^q) = \frac{C}{D}\left(\E[\tilde{r}^p] - \frac{A}{C}\right)\left(\E[\tilde{r}^q] - \frac{A}{C}\right) + \frac{1}{C} \\
\sigma^2(w^p) = \frac{1}{D}\left(C\E[\tilde{r}^p]^2 - 2A\E[\tilde{r}^p] + B\right).
\end{align*}
\item The PF is a hyperbola in the mean-standard-deviation (MSD) diagram, and a parabola in the mean-variance (MVAR) diagram. The portfolio with the minimal standard deviation $w^MVP$ defines the \textit{minimum variance point} in the diagrams,
\[
w^{MVP} = \frac{V^{-1} \mathbb{I}}{C}.
\]
\item For every expected return, there is one point on the PF. For every standard deviation or variance (beyond the minimal variance point) there are two points on the PF. We define the \textit{efficient portfolio frontier} (EPF) as the part of the PF that is above the minimal variance point, i.e.
\[
\text{EPF} = \left\{w \in PF : \E[\tilde{r}] > \frac{A}{C}\right\}.
\]
The lower branch is the inefficient part of the PF. In case agents maximize their expected utility, then restriction to the EPF is justified if the utility function is increasing in average return and decreasing in variance.
\item Given a portfolio $w^p$ belonging to the PF, there exists a portfolio $w^{zc(p)}$ (also belonging to the PF) that has zero covariance with $w^p$. For the return of this portfolio we find
\[
\E[\tilde{r}^{zc(p)}] = \frac{A}{C} - \frac{\frac{D}{C^2}}{\E[\tilde{r}^p] - \frac{A}{C}}
\]
\end{enumerate}
\begin{proposition} \label{prop:3.4}
Let $w^q$ be a portfolio and let $w^p$ be a portfolio that belongs to the PF, and that is not the minimal variance portfolio. Then
\begin{align*}
w^q \in PF &\implies \tilde{r}^q = (1-\beta_{qp})\tilde{r}^{zc(p)} + \beta_{qp}\tilde{r}^p, \\
w^q \not\in PF &\implies \tilde{r}^q = (1-\beta_{qp})\tilde{r}^{zc(p)} + \beta_{qp}\tilde{r}^p + \tilde{\epsilon}^{qp},
\end{align*}
where
\[
\E[\tilde{\epsilon}^{qp}] = 0, \;\;\;\;\; \beta_{qp} = \frac{cov(w^p, w^q)}{\sigma^2(w^p)}.
\]
\end{proposition}
Note that the return of any portfolio is the linear combination of a portfolio on the frontier, its zero-covariance counterpart and a random variable with zero mean and zero covariance with the frontier portfolios. Furthermore, the coefficients of the frontier portfolio and its zero covariance counterpart are the regression coefficients of the portfolio on those frontier portfolios.
\section{Portfolio Frontier with a Risk-Free Asset}
Consider now the setup of $N$ risky assets, with a risk-free asset included. The constraint $W \in \Delta_N$, with $\Delta_N = \{x \in \R^N : \sum_{n=1}^{N} = 1\}$. We can relax this by investing the fraction $1 - w^\top \mathbb{I} = 1 - \sum_{n=1}^{N} w_n$ in the risk-free asset. In order to identify the portfolio frontier that includes the risk-free asset (PF\textsuperscript{*}), we have to solve the same problem, but under a single constraint:
\begin{align*}
\min_w \frac{1}{2}w^\top V w, \;\;\; w \in \R^N, \\
\text{s.t.} \;\;\; w^\top e + (1-w^\top \mathbb{I})r_f = \E[\tilde{r}^p].
\end{align*}
As the PF is a hyperbola in the MSD diagram, we have that, if $r_f \not= \frac{A}{C}$, then there is exactly one point in which the PF and PF\textsuperscript{*} coincide. If $r_f < \frac{A}{C}$, this is a point on the EPF. If $r_f > \frac{A}{C}$, then this is a point on the lower branch of the PF, so it is inefficient.
\begin{figure}[ht]
\centering
\incfig{epf}
\caption{The efficient portfolio frontier.}
\label{fig:epf}
\end{figure}
\begin{proposition}
Let $w^q$ be a portfolio and let $w^p$ be a portfolio that belongs to the PF\textsuperscript{*}, where the expected return of $w^p$ is not equal to the risk-free rate. Then
\begin{align*}
w^q \in PF^* &\implies \tilde{r}^q = (1-\beta_{qp})r_f + \beta_{qp}\tilde{r}^p, \\
w^q \not\in PF^* &\implies \tilde{r}^q = (1-\beta_{qp})r_f + \beta_{qp}\tilde{r}^p + \tilde{\epsilon}^{qp},
\end{align*}
where
\[
cov(\tilde{r}^q, \tilde{\epsilon}^{qp}) = \E[\tilde{\epsilon}^{qp}] = 0, \;\;\;\;\; \beta_{qp} = \frac{cov(\tilde{r}^q, \tilde{r}^p)}{\sigma^2(\tilde{r}^p)}.
\]
Moreover, for any portfolio $w^q$,
\[
\E[\tilde{r}^q] - r_f = \beta_{qp}(\E[\tilde{r}^p] - r_f).
\]
\end{proposition}
That brings us to the \textit{mutual fund separation} property: For any portfolio $w^q$, there is a portfolio that is on the EPF\textsuperscript{*} which is a linear combination of the risk-free and tangent portfolio on the EPF, that is better in the SSD sense.
\section{CAPM}
Consider an economy with $I$ agents ($i = 1, \dots, I$). The wealth of agent i at time $t=0$ is given by $W^i_0 \geq 0$. The percentage of wealth that agent $i$ holds in asset $n$ is given by $w^i_n$, with $n = 1, \dots, N$. The return of asset $n$ is given by the random variable $\tilde{r}^n$, with finite variance. The wealth of the economy is then given by
\[
W_0^m = \sum_{i=1}^{I} W_0^i.
\]
The percentage of wealth that the economy holds in asset $n$ is given by
$w^m_n$ and $w^m$ is called the market portfolio. We have the following assumptions:
\begin{itemize}
\item Perfectly competitive markets.
\item No transaction costs and taxes, and assets are perfectly divisible.
\item Agent's preferences can be represented through the expected utility.
\item Agents are risk averse.
\item The probabilistic framework is common knowledge.
\item The mutual fund separation property holds, or agent's optimal portfolios belong to the PF.
\item The economy is closed.
\end{itemize}
In equilibrium, the market demand of an asset must be equal to its supply. That is,
\[
w^m_n = \sum_{i=1}^{I} w^i_n \frac{W^i_0}{W^m_0}
\]
\begin{remark}
In practice this means that we can 'read' the market portfolio from the market using the \textit{market captialization weights}, i.e. $w^m_n = (\# \text{shares outstanding} \times \text{share price}) / \text{total value of the economy}$.
\end{remark}
If agents' portfolios belong to the PF, then the market portfolio $w^m$ also belongs to the PF. If $w^m$ is different from the minimal variance point, we can use \cref{prop:3.4} to obtain
\begin{align*}
\E[\tilde{r}^q] = \E[\tilde{r}^{zc(m)}] + \beta_{qm}\left(\E[\tilde{r}^m] - \E[\tilde{r}^{zc(m)}]\right), \\
\tilde{r}^m = \sum_{n=1}^{N} w^m_n \tilde{r}^n, \;\;\; \beta_{qm} = \frac{cov(w^q, w^m)}{\sigma^2(w^m)}.
\end{align*}
This is a straight line in the beta-expected return plane. The line is called the \textit{security market line}. If the slope of the line is positive, then the market portfolio is efficient. If the slope is negative, then the market portfolio is inefficient.
Including the risk free asset gives the relationship
\[
\E[\tilde{r}^q] = r_f + \beta_{qm}\left(\E[\tilde{r}^m] - r_f\right).
\]
If $r_f < \frac{A}{C}$ then the market portfolio is efficient, and all agents hold a combination of the risk-free asset and the market portfolio. If $r_f > \frac{A}{C}$, then the market portfolio is inefficient, and all agents hold only the risk-free asset.
Conclusions:
\begin{itemize}
\item The CAPM is a one-factor model, where beta is the factor that linearly relates the risk premium of a portfolio.
\item The risk premium of a portfolio has the same sign as the risk premium of the market portfolio in case of positive correlation. In case of negative correlation, it has the opposite sign.
\item The return of a portfolio can be decomposed into a constant, a risk component associated with market return (systemic risk), and a component with zero mean and zero correlation with market return (idiosyncratic risk).
\end{itemize}
\section{Empirical Test CAPM}
To test CAPM on historical data we have to estimate beta and the expected returns. We can write the asset returns as:
\[
\tilde{r}_n(t) = \E[\tilde{r}_n] + \beta_{nm}(\tilde{r}^m(t) - \E[\tilde{r}^m]) + \tilde{\epsilon}_n(t)
\]
Now we impose CAPM restrictions and get:
\[
\tilde{r}_n(t) - r_f = \alpha_n + \gamma_{n}(\tilde{r}^m(t) - r_f) + \tilde{\epsilon}(t)
\]
We assume that the asset returns are stationary, so this models holds period by period. There is no observarble market portolio, so we take a stock index. Then we can estimate $\beta$ and test the hypothesis on the daily asset returns.
The result is that in the early 70's, much evidence for the CAPM was found. Thereafter, additional explantory variables are introduced, e.g. size of the company, leverage, value effect, momentum effect, etc. There are three groups: loyalists, who think that counterevidence is a statistical flaw; revisionists, who think that extra risk factors need to be added; heretics, who think the market is irrational.
\lecture{9}
\section{Arbitrage Pricing Theory}
If we take two other assumptions to derive the risk premia, we can derive a different model. Firstly, there are no arbitrage opportunities in the market. Secondly, a multifactor model holds for the asset returns, and the model is common knowledge.
Suppose we have $N$ risky assets, traded in the market in a two period model $(t=0, t=1$). The returns over this period are given by $K$ risk factors:
\[
\tilde{r}_n = e_n + \sum_{k=1}^{K} b_{nk} \tilde{\delta}_k, \;\;\; n = 1, \dots, N.
\]
Here, $\tilde{\delta}_k$ are the risk factors. We assume that $\E[\tilde{\delta}_k] = 0$, and therefore the expected return of asset $n$ is given by $e_n$. That is, $\tilde{r} = e + B\tilde{\delta}$, written in vector form.
\begin{proposition}[Proposition 5.2] \label{prop:5.2}
Given the linear multifactor model for asset returns. Without arbitrage opportunities in the market, there exists a vector $\lambda \in \R^K$ such that:
\[
e - r_f\mathbb{I} = B\lambda.
\]
\end{proposition}
\begin{proof}
For any portfolio $w \in \R^N$ (representing the proportions of wealth invested in the $N$ risky assets, thereby normalizing the initial wealth to 1), the associated return is given by $w^\top \tilde{r} = w^\top \E[\tilde{r}] + w^\top B \tilde{f}$. Considering a portfolio $w$ such that $w^top \mathbb{I} = 0$, and $w^\top B = 0$, the absence of arbitrage implies that $w^\top \E[\tilde{r}] = 0$ (otherwise, it would be possible to create a portfolio with zero cost and strictly positive return). Equivalently, this means that $\ker(B^\top) \cap \ker(\mathbb{I}^\top) \subseteq \ker(\E[\tilde{r}]^\top)$. Hence, there exists a couple $(\lambda, \lambda_0) \in \R^{K+1}$ such that
\[
\E\tilde{r} = \mathbb{I}\lambda_0 + B\lambda.
\]
If the kernel of $B^{\top}$ is a subset of the kernel of $\mathbb{I}^{\top}$, then it holds that $\mathbb{I}=B \gamma$, for some $\gamma \in \mathbb{R}^K$. In this case, the last relation implies that
\[
\mathbb{E}[\tilde{r}]=B\left(\lambda_0 \gamma+\lambda\right)=B\left(\lambda_0 \gamma-r_f \gamma+\lambda\right)+r_f \mathbb{I},
\]
thus proving \cref{prop:5.2}. Suppose now that the kernel of $B^{\top}$ is not a subset of the kernel of $\mathbb{I}^{\top}$, so that there exists a vector $w \in \mathbb{R}^N$ such that $w^{\top} B=0$ but $w^{\top} \mathbb{I} \neq 0$. We claim that, due to the absence of arbitrage opportunities, it holds that $\lambda_0=r_f$. Indeed, suppose on the contrary that $\lambda_0>r_f$ (the case $\lambda_0<r_f$ can be treated in an analogous way). Let $\bar{w}:=w /\left(w^{\top} \mathbb{I}\right)$ and consider the following strategy: borrow one unit of wealth at the risk free rate $r_f$ and invest that unit of wealth in the $N$ risky assets according to the portfolio $\bar{w}$. This strategy has zero cost at time $t=0$ and delivers a return of $\bar{w}^{\top} \mathbb{I}\left(\lambda_0-r_f\right)=\lambda_0-r_f>0$ at time $t=1$, thus contradicting the absence of arbitrage opportunities.
\end{proof}
The expected return of asset $n$ is then given by
\[
\E[\tilde{r}_n] = e_n + \sum_{k=1}^{K} b_{nk} \lambda_k, \;\;\; n = 1, \dots, N.
\]
However, this does not take idiosyncratic risk into account. Introducing that gives
\[
\tilde{r}_n = e_n + \sum_{k=1}^{K} b_{nk} \tilde{\delta}_k + \tilde{\epsilon}_n, \;\;\; n = 1, \dots, N,
\]
where $\E[\tilde{\delta}_k] = 0$, $\E[\tilde{\epsilon}_n] = 0$.
An \textit{asymptotic arbitrage} is a subsequence of portfolios such that $w^{N'T}\mathbb{I} = 0$, and
\[
\lim_{N' \to \infty} \E[w^{N'T}\tilde{r}^{N'}] = +\infty \;\;\; \text{and} \;\;\; \lim_{N' \to \infty} \sigma^2(w^{N'T}\tilde{r}^{N'}) = 0.
\]
\begin{proposition}
Given the linear multifactor model for asset returns, without asymptotic arbitrage. There exists a scalar $\lambda^N_0$, a vector $\lambda^N \in \R^K$ and a positive constant $\alpha$ such that:
\[
\sum_{n=1}^{N}\left(e^N_n - \lambda^N_n - \sum_{k=1}^{K} b^N_{nk} \lambda^N_k\right)^2 \leq \alpha.
\]
\end{proposition}
If we now add the risk-free asset to the model, we get that $\lambda^N_0 = r_f$. The consequence is
\[
e^N - \mathbb{I}r_f \approx B^N \lambda^N.
\]
\lecture{10}
\section{Fundamental Theorem of Asset Pricing}
Assume that we have $S$ states of the world, $s = 1, \dots, S$ with the probability that we are in state $s$ given by $\pi_s$. The economy has $N$ assets with initial price vector $q \in \R^N$, and the portfolio weights $w \in \R^N$, so the initial value is given by $q^\top w$. The matrix $D \in \R^{S \times N}$ contains the dividends of the $N$ assets at time $t=1$ in each state of the world. So, $Dw$ is the payoff in $S$ states of the world at time $t=1$ of the portfolio.
Given the price-dividend couple $(q, D)$, a portfolio $w$ constitutes an arbitrage if one of the following holds:
\begin{itemize}
\item Arbitrage of the first type: $q^\top w \leq 0$ and $Dw > 0$.
\item Arbitrage of the second type: $q^\top w < 0$ and $Dw \geq 0$.
\end{itemize}
From the no-arbitrage assumption, two properties arise. The \textit{law of one price} states that two portfolios with the same dividend in every state of the world at $t=1$, should have the same market value at $t=0$. \textit{Arbitrage without risk} means that there is no portfolio with non-positive market value at $t=0$ getting a constant positive amount of money at $t=1$.
We define the pricing functional as the function $q : \R^S \mapsto \R$ that maps a pay-off (or consumption plan) into the initial market value. The value $q(1_s)$ is the price of reaching state of the world $s$ (cf. the Arrow-Debreu price). The pricing functional is linear if $q(\alpha c_1 + \beta c_2) = \alpha q(c_1) + \beta q(c_2)$ for all $x, y \in \R^S$ and $\alpha, \beta \in \R$. There exists a linear pricing functional if and only if there exists a solution to the system $q = D^\top m$, with $m_s = q(\mathbb{I}_s)$. That is,
\[
q_n = \sum_{s=1}^{S}d_{ns}m_s,
\]
and $q(c) = m^\top c$.
The pricing functional is strictly positive if $m$ is strictly positive. If now matrix $D$ has full rank and $N \geq S$ (so the market is complete) then the state price vector $m$ exists and is unique. Otherwise markets are incomplete and the state price vector is not unique.
\begin{proposition}[Fundamental Theorem of Asset Pricing]
Given a couple of of a price vector $q \in \R^N$ and a dividend matrix $D$, there are no arbitrage opportunities in the market if and only if there exists a strictly positive linear pricing functional.
\end{proposition}
A strictly positive pricing functional has an interesting probabilistic interpretation. If we divide $q_n$ by $m_0 = \sum_{s=1}^{S}m_s$, we get
\[
\frac{q_n}{m_0} = \sum_{s=1}^{S} d_{ns}\pi^*_s, \;\;\; \pi^*_s = \frac{m_s}{m_0}.
\]
Here, $(\pi_1^*, \dots, \pi_S^*)$ constitutes a probability measure that is equivalent to the original measure $\pi$. Now we can write $q_n = m_0 \E^*[\tilde{d}_n]$. Thus, $m_0$ is the price of receiving 1 in every state of the world at time $t=1$. If there is a risk-free asset, then the no arbitrage condition should yield that $m_0 = \frac{1}{1+r_f}$. Thus, $q_n = \frac{1}{1 + r_f} \E^*[\tilde{d}_n]$. The result is that $\E^*[\tilde{r}_n] = r_f$.
The strictly positive linear pricing functional defines the risk-neutral probability space. So, if the maket is not complete, there exist many pricing functionals and as a result thhere exist many risk-neutral probability measures. Therefore, completeness of the market relates to uniqueness of the risk-neutral measure.
\end{document}