An approximate solution for the power optimization under predictable returns

Dmytro Ivasiuk

Department of Statistics, European University Viadrina, Frankfurt(Oder), Germany

E-mail: [email protected]

Abstract This work presents an approximate single period solution of the portfolio choice problem for the investor with a power utility function and the predictable returns. Assuming that gross portfolio returns approximately follow a log-normal distribu- tion, it was possible to derive the optimal weights. As it is known, the log-normal distribution seems to be a good proxy of the normal distribution in case if the stan- dard deviation of the last one is way much smaller than its mean. Thus, succeeding to model the behaviour of asset returns as a multivariate normal (conditional nor- mal) distribution (what is a popular way to proceed in academic literature) would be the case to apply the ideas discussed in this work. Besides, the paper provides a simulation study comparing the derived result to the well-know numerical solution.

Keywords: CRRA, power utility, utility optimization, numerical solution, log-normal distribution.

1 Introduction

The current paper discusses portfolio optimization as the utility of wealth maximization. The classical problem stands for deriving the value function given by the whole investment period with the use of the Bellman backward method (see Pennacchi, 2008; Brandt, 2009). One can find it as the typical solution for the investment strategies for different types of utility functions (see Bodnar et al., 2015a,b). However, most of the time, it is hard

arXiv:1911.06552v3 [q-fin.PM] 21 Jan 2021 to succeed and to derive an analytical solution. Because of the difficulty to perform the calculations (see Bodnar et al., 2015b; Campbell and Viceira, 2002), it is needed to have some specific assumptions. On the other hand, one can always use numerical and approximative results (see Brandt et al., 2005; Broadie and Shen, 2017). However, the quadratic utility has a closed-form multi-period solution which requires to estimate a variance- and expected value of the returns (see Bodnar et al., 2015a). W 1−γ This work deals with the power utility of wealth function U(W ) = 1−γ , γ > 0, which classifies as a constant relative aversion (CRRA) utility. The interpretation of the CRRA invented by John W. Pratt stands for the person who’s investing decisions do not rely on his or her initial wealth (see Campbell and Viceira, 2002; Pennacchi, 2008), which is, in fact, observable in findings of the paper.

1 To derive the optimal weights, we will consider that asset returns follow a multivariate normal distribution. The similar way to proceed is to use a conditional multivariate normal distribution, which appears, for instance, in the vector-autoregressive process given multivariate normal error terms. Those models have become a popular choice to model the return path in recent literature (see Bodnar et al., 2015a,b; Barberis, 2000; Brandt and Santa-Clara, 2006). Consequently, it comes up with the multivariate normal asset returns and, the resulting portfolio returns will be normally distributed, as a linear combination of normal variables. However, it gives no benefits in case of the power utility function. It is convenient to have the log-normal distribution for the gross portfolio returns, to calculate the expected value of a power function of wealth. For that reason, the log-normal distribution emerges as an approximation of the normal distribution. Precisely speaking, the idea is to model the returns using the proxy distribution which behaves similar to the original one. Note that the log-normal distribution often used to model stock returns in literature. However, the resulting portfolio returns do not remain log-normal. For that reason, the idea is to switch to the limit case and work in the continuous-time setting (see Campbell and Viceira, 2002; C¸etinkaya and Thiele, 2016; Herzel and Nicolosi, 2019). The main result of the paper provides approximate single-period optimal portfolio weights for the power utility function. The approximation is based on a similar behaviour of the normal N(µ, σ2) and log-normal LN(ln µ, σ2/µ2) densities if σ/µ approaches zero (see Taras et al., 2020). For instance, let µ = 1 and σ = 0.04 then both densities look pretty similar, as presented in Figure 1. Indeed, testing simulated samples from the corresponding log-normal distribution would frequently fail to reject the normality hypothesis. Performing Shapiro-Wilk test for 105 samples of 2500 values simulated from log-normal distribution LN(0, 0.042) has a p-value greater than 0.05 in around 43% cases. A similar situation holds in practice at a stock market trading: the gross returns fluctuate around value zero within a small interval. It is also notable that the correct definition of the approximation requires a positive definite µ. That is itself a fair assumption since nobody expects to lose all the invested money. One might also question the normality setup because it implies a possibility of negative wealth, although it is quite small from a practical point of view (x ∼ N(1, 0.1) ⇒ p(x < 0) = 7.62·10−24). On the other hand, this issue vanishes under the log-normality (because of the positive definite random variables), that makes it so attractive for the power utility function. One should also clarify that the approach presented in this paper is slightly different from Taras et al. (2020). The present work has the normal distribution to model returns, whereas in Taras et al. (2020) the log-normal distribution has the same mean and variance as gross portfolio returns. As it is known, the power utility optimization has no analytical solution and therefore, for an empirical study we will compare the derived result to the well known numerical solution (see Brandt et al., 2005), since both of them are an approximation. In general, the preferable among two portfolios is the one with a higher expected utility of the final wealth. Thus it is done in the same manner. Simulating the return path from a vector autoregressive model with normally distributed error terms, one can calculate the final utility of wealth for both strategies and examine the outcomes. Even though the numerical solution provides multi-period weights, the derived result (built under a single-period setup) outperforms. Simulations have shown, that the numerical wights produce a smaller sample mean for the final utility of wealth, which means, that in case of a normally distributed returns the derived solution dominates.

2

10 N(1, 0.0016) lnN(0, 0.0016) 8 6 4 2 0

0.8 0.9 1.0 1.1 1.2

Figure 1: Normal and Log-normal densities

2 Framework

Below one can find a framework used to present the portfolio selection problem for a power utility function. Let r denote the one-period random return vector on risky assets and let rf be the return on the risk-free asset. Let ω define portfolio weights of an initial wealth W0 allocated between risky assets. Therefore, if Rf = 1 + rf is the gross return on the risk-free asset and R = r − rf 1 is the vector of excess return on the risky assets, 0 then the investor’s wealth at the end of investment period is W = W0 (Rf + ω R). So it is considered that in investor distributes all the money between risky assets and one risk-free asset without any consumption. Also let us define the mean vector and the variance-covariance matrix of the excess returns with the next way: µ = E[r − rf 1], Σ = V ar[r − rf 1]. The purpose of the paper is to derive the optimal portfolio weights in order to maximise the expected utility of the final wealth: ω∗ = arg maxE [U (W )] . (1) {ω} Precisely speaking, this is a single-period setup of the utility maximization problem (see, Brandt and Santa-Clara, 2006; Pennacchi, 2008; Bodnar et al., 2015a). The analyzed utility U(·) is a power function of wealth W 1−γ U(W ) = , γ > 0, γ 6= 1, (2) 1 − γ which has constant relative (CRRA) γ and belongs to HARA family (see C¸anako˘gluand Ozekici,¨ 2010, 2012).

3 To model the behavior of excess returns, we will use the multivariate normal distribu- tion, N(µ, Σ). On the other hand, to approximate the resulting normal distribution of 0 0 the gross portfolio returns, N(ω µ + Rf , ω Σω), we choose the log-normal distribution, 0 ω0Σω2 LN(ln(ω µ+Rf ), 0 2 ) (see Taras et al., 2020). One can also consider the conditional (ω µ+Rf ) normal distribution which is a case when we talk about vector autoregressive processes (see, Barberis, 2000; Campbell and Viceira, 2002; Bodnar et al., 2015b). For instance, if the excess returns follow the VAR process of order one

Rt = ϕ + ΦRt−1 + εt, εt ∼ N (0, Σ) , 0 then the gross portfolio returns Rf + ω Rt will have the conditional normal distribution 0 0 with the mean Rf + ω (ϕ + ΦRt−1) and the variance ω Σω. Precisely speaking, only the normality of error terms is crucial, since the idea behind is to approximate the normal distribution with the log-normal based on the convenience to calculate the expected value of a power function of a log-normal random variable:  1  E Xλ = exp αλ + β2λ2 , if X ∼ LN(α, β2), λ ∈ . (3) 2 R

3 Approximate solution of the power utility opti- mization

The main finding of this paper presents an approximate solution for a single-period port- folio optimization problem (1). We also consider an investor with a power utility of wealth function (2) and the excess returns to follow the multivariate normal distribution. Particularly any model with a multivariate normal (conditional) distribution satisfies the derived solution. Theorem 1. Assume that the excess returns R follow the multivariate normal distribution with the covariance matrix Σ and the mean vector µ. If γ ≥ 1 + 4µ0Σ−1µ, (4) then the approximate solution of the optimization problem (1) with the power utility func- tion (2) is given by  2 γ − 1 − p(γ − 1)2 − 4(γ − 1)µ0Σ−1µ ∗ −1 ω = Σ µRf . (5) 4(γ − 1) µ0Σ−1µ2 The proof of the theorem one can find in the Appendix section. Note that optimal weights (5) do not exist for all the values γ. However, having a minimum level of relative risk aversion can be explained by the fact, that considering the log-normal distribution we, thereby, bound the gross portfolio returns from bellow. Thus, it means that an investor rejects a risky trade where it is possible to lose everything on the market. Moreover, the next section provides empirical results for the lower bound of γ presented as a histogram, which gives the idea of its possible location. Corollary 1. Given by Theorem 1, the corresponding portfolio has its expected excess return and variance as a parabola: (ω∗0µ)2 = ω∗0Σω. µ0Σ−1µ

4 Corollary 2. Given by Theorem 1, the corresponding portfolio has its expected excess return and variance being decrease functions of the coefficient of relative risk aversion γ.

The first statement is obvious and for Corollary 2 see proof in the Appendix section. Both corollaries describe two economic insights. The first one is similar to the mean- variance efficient frontier trade-off (see, Merton, 1972; Pennacchi, 2008; Brandt, 2009); namely, the higher expected return is achievable only by taking more risk, and vice versa. The second one corresponds to the interpretation of risk aversion, that is having large γ means to aim for the less volatility in exchange for a portion of wealth.

Remark 1. If γ = 1 + 4J the optimal weights in Theorem 1 will be

Σ−1µ ω∗ = R , 0 f µ0Σ−1µ

0 ∗ with the resulting portfolio excess return µ ω0 = Rf . Corollary 3. Let Theorem 1 holds. If we consider no investment into the risky asset, then the optimal weights will look as follows:

Σ−1µ ω = . tgc 10Σ−1µ The corresponding coefficient of relative risk aversion is given by:

µ0Σ−1µ 2 10Σµ γtgc = 0 −1 + Rf + 1. 1 Σ µ Rf See the proof of Corollary 3 in the Appendix section. As one might see, the result of Theorem 1 has got an interesting property. Assuming to allocate the wealth only to the risky assets, it will produce the same weights as the the tangency portfolio has in the mean-variance paradigm of Markowitz (1952) (see Brandt, 2009).

4 Comparison study

The empirical study compares the approximate weights derived in Theorem 1 to the well known numerical result introduced by Brandt et al. (2005). The last one is a universal solution obtainable with a Taylor series expansion of the value function, possible to use for any given utility. However, it is a time consuming iterative procedure, since one should run a large number of simulations to achieve high accuracy. Presented below, one can find the numerical solution as the fourth-order expansion suggested by Brandt et al. (2005):

5    −1  2 ˆ T −1  ∂ U WT  Y 0 2 0  ωt(i + 1) ≈ − Et ωˆ Rs+1 + Rf Rt+1R Wt  ˆ 2 s t+1  ∂WT s=t+1    (  ˆ T −1  ∂U WT Y × E ωˆ 0 R + R  R t  ˆ s s+1 f t+1 ∂WT s=t+1   (6)  3 ˆ T −1  ∂ U WT 1 Y 0 3 0 2 2 + Et  ωˆ sRs+1 + Rf (ωt(i)Rt+1) Rt+1 Wt 2 ˆ 3 ∂WT s=t+1    4 ˆ T −1  ) ∂ U WT 1 Y 0 4 0 3 3 + Et  ωˆ sRs+1 + Rf (ωt(i)Rt+1) Rt+1 Wt , 6 ˆ 4 ∂WT s=t+1 where ωˆ s defines already calculated weights (s = t + 1,...,T − 1) and T −1 ˆ Y 0  WT = WtRf ωˆ sRs+1 + Rf . (7) s=t+1 ˆ After calculating the derivatives of the power utility and substituting WT , one has:

( " T −1 #)−1 1 Y 1−γ ω (i + 1) ≈ E ωˆ 0 R + R  R R0 t γ t s s+1 f t+1 t+1 s=t+1 ( " T −1 # Y 0 1−γ × Rf Et ωˆ sRs+1 + Rf Rt+1 s=t+1 (8) " T −1 # γ(γ + 1) Y 1−γ 2 + E ωˆ 0 R + R  (ω0 (i)R ) R 2R t s s+1 f t t+1 t+1 f s=t+1 " T −1 #) γ(γ + 1)(γ + 2) Y 1−γ 3 − E ωˆ 0 R + R  (ω0 (i)R ) R , 6R2 t s s+1 f t t+1 t+1 f s=t+1 where ( " T −1 #)−1 1 Y 1−γ ω (0) ≈ E ωˆ 0 R + R  R R0 t γ t s s+1 f t+1 t+1 s=t+1 (9) " T −1 # Y 0 1−γ × Rf Et ωˆ sRs+1 + Rf Rt+1 s=t+1 and 1 ω (i + 1) ≈ {E [R R0 ]}−1 T −1 γ T −1 T T ( γ(γ + 1) h 0 2 i × Rf ET −1 [RT ] + ET −1 ωT −1(i)RT RT (10) 2Rf ) γ(γ + 1)(γ + 2) h 0 3 i − 2 ET −1 ωT −1(i)RT RT , 6Rf

6 with R ω (0) ≈ f {E [R R0 ]}−1 E [R ] . (11) T −1 γ T −1 T T T −1 T Hence, this is a complete backward scheme on how to obtain the numerical solution of a multi-period investment strategy similar to the Bellman method. Like it was mentioned before, there are no wealth components in the weights formula because of the constant relative risk aversion utility function. The only task left is to calculate the expectations where the law of large numbers is applied (see Barberis, 2000): if f(x) defines a probability density function of a random variable x, then the expectation E [g(x)] can be estimated by N 1 X g(y ) N i i=1 N where {yi}i=1 is a large sample from the distribution given by f(x). For instance, to do all the calculations N = 105 and i = 20 was taken, and for simplicity, we use a constant return of 0.01 on the risk-free asset. To simulate the excess returns, we take a model for a stock, a bond, and a state variable, suggested by Brandt and Santa-Clara (2006):

 s       s    ln 1 + rt+1 0.0059 0.006 εt+1 Rt b  b = ln 1 + rt+1  =  0.0007  + 0.0053 × zt + εt+1 , (12) zt+1 z zt+1 −0.0028 0.9597 εt+1 with  s     εt+1 0.0018 0.0002 −0.0005 b εt+1 ∼ MVN 0,  0.0002 0.0006 0.0007  . (13) z εt+1 −0.0005 0.0007 0.0802 According to Theorem 1, the mean vector is

0.0059  0.006  µ = + × z 0.0007 0.0053 t and the covariance matrix is given by

0.0018 0.0002 Σ = , 0.0002 0.0006 which are used in the comparison study. Table 1 presents sample means of the power utility of a final wealth for several combi- nations of γ and T for the numerical (Numerical) solution and the portfolio weights from Theorem 1 (Approximate). Although the expression (5) provides the one-period con- straint, one can see that the corresponding values are very close, which means that the derived result shows a high level of approximation which even outperforms the numerical one. Besides, Figure 2 - 5 describe the behaviour of the empirical cumulative distribu- tion function of both strategies for several combinations of the coefficient of relative risk aversion γ and the investment range T . Along with the comparison study, we demonstrate some properties of the result derived in Theorem 1 and illustrate the ideas described in Corollary 1 and Corollary 2. The

7 T γ 5 10 15 20 Method -0.18893 -0.06217 -0.02953 -0.01623 Numerical 6 -0.18679 -0.06122 -0.02901 -0.01596 Approximate -0.14346 -0.03482 -0.01261 -0.00502 Numerical 12 -0.13735 -0.03283 -0.01154 -0.00474 Approximate -0.10872 -0.01957 -0.00519 -0.00152 Numerical 18 -0.09965 -0.01738 -0.00451 -0.00135 Approximate -0.08294 -0.01103 -0.00216 -0.00047 Numerical 24 -0.07191 -0.00925 -0.00183 -0.00039 Approximate

Table 1: The sample means of the power utility of the final wealth for the numerical solution and the derived strategy for several combinations of γ and T

γ=5, T=6 γ=5, T=12

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.5 −0.4 −0.3 −0.2 −0.1 −0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0.0

γ=5, T=18 γ=5, T=24

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.6 −0.5 −0.4 −0.3 −0.2 −0.1 0.0 −0.8 −0.6 −0.4 −0.2 0.0

Figure 2: Empirical distribution function of the power utility for the numerical and the derived strategies for several combinations of γ and T

8 γ=10, T=6 γ=10, T=12

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.15 −0.10 −0.05 −0.15 −0.10 −0.05 0.00

γ=10, T=18 γ=10, T=24

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.15 −0.10 −0.05 0.00 −0.15 −0.10 −0.05 0.00

Figure 3: Empirical distribution function of the power utility for the numerical and the derived strategies for several combinations of γ and T

γ=15, T=6 γ=15, T=12

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.10 −0.08 −0.06 −0.04 −0.02 0.00 −0.06 −0.04 −0.02 0.00

γ=15, T=18 γ=15, T=24

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.06 −0.04 −0.02 0.00 −0.020 −0.015 −0.010 −0.005 0.000

Figure 4: Empirical distribution function of the power utility for the numerical and the derived strategies for several combinations of γ and T

9 γ=20, T=6 γ=20, T=12

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.04 −0.03 −0.02 −0.01 −0.04 −0.03 −0.02 −0.01 0.00

γ=20, T=18 γ=20, T=24

1.0 1.0 Numeric Numeric Approximate Approximate 0.8 0.8

0.6 0.6 Fn(x) Fn(x) 0.4 0.4

0.2 0.2

0.0 0.0

−0.014 −0.012 −0.010 −0.008 −0.006 −0.004 −0.002 0.000 −0.006 −0.004 −0.002 0.000

Figure 5: Empirical distribution function of the power utility for the numerical and the derived strategies for several combinations of γ and T

Mean−variance trade−off

γ=2.1 0.015

γ=2.5

0.010 γ=3

γtgc=3.93

Expected portfolio excess returnExpected portfolio excess γ 0.005 =5

0.000 0.005 0.010 0.015

Portfolio return variance

Figure 6: Mean-variance trade-off of the portfolio derived in Theorem 1

10 Expected portfolio excess return as a function of γ 0.015 0.010 Expected portfolio excess returnExpected portfolio excess 0.005

2 4 6 8 10

Coefficient of relative risk aversion, γ

Figure 7: Expected portfolio excess return as a function of γ

Portfolio variance as a function of γ 0.015 0.010 Portfolio variance Portfolio 0.005 0.000 2 4 6 8 10

Coefficient of relative risk aversion, γ

Figure 8: Portfolio variance as a function of γ

11 4

3

2 Density

1

0

1.0 1.5 2.0 2.5 3.0 3.5 4.0

γmin

Figure 9: Histogram of the lower bound of γ-values in Theorem 1

0.005483 hypothetical mean vector and the covariance matrix of excess returns are 0.000457 0.0018 0.0002 and respectively. In Figure 6, one can find the parabola plot of the 0.0002 0.0006 expected portfolio excess returns with the corresponding variances. Also, we added several points representing values of the coefficient of relative risk aversion, including the tangency portfolio. One can observe that those dots move towards the edge of the parabola as γ increases. This particular behaviour of the portfolio mean and variance being a decrease function of the coefficient of relative risk aversion is also well described in Figure 7 and Figure 8. The last part of the simulation study is a visualization of the lower bound of the coefficient γ in Theorem 1. Figure 9 presents the histogram of the expression 1 + µ0Σ−1µ estimated form 105 samples given by the model (12). One can observe that most of the values are between one and two. However, there are some points outside of the interval, not exceeding value four. Meanwhile, the literature provides the median value of relative risk aversion being approximately seven (see Pennacchi, 2008).

5 Appendix

0 Lemma 1. Assume that the portfolio gross return Rf + ω R at the end of the investment period is normally distributed, i.e.

0 0 0 Rf + ω R ∼ N(Rf + ω µ, ω Σω).

12 √ 0 ω0Σω Let Rf + ω µ > 0 and R +ω0µ → 0. f   0 (approx.) 0 ω0Σω Then R + ω R ∼ LN ln (R + ω µ) , 2 f f 0 (Rf +ω µ) Proof of Lemma 1. According to Taras et al. (2020), assuming that µ > 0 and σ/µ → 2  σ2  0, the difference between the distribution functions of N (µ, σ ) and LN ln µ, µ2 ap- proaches zero. Proof of Theorem 1. For the power utility function, one has " # W 1−γ (R + ω0R)1−γ W 1−γ h i E [U(W )] = E 0 f = 0 E (R + ω0R)1−γ . 1 − γ 1 − γ f

Assuming that R ∼ N (µ, Σ), given by Lemma 1 it holds:

1−γ " 2 0 # W0 0 (1 − γ) ω Σω E [U(W )] ≈ exp (1 − γ) ln (Rf + ω µ) + . (14) 0 2 1 − γ 2 (Rf + ω µ)

In order to maximize the expected utility (14) it is required to derive the first order conditions (FOCs) with respect to ω. The FOCs are: " 0 # ∂ 0 1 − γ ω Σω ln (Rf + ω µ) + = 0 (15) 0 2 ∂ω 2 (Rf + ω µ) The partial derivation leads to:

0 2 0 0 µ Σω (Rf + ω µ) − ω Σω(Rf + ω µ)µ 0 + (1 − γ) 0 4 = 0, ω Rf + µ (Rf + ω µ) or 0 0 Σω (Rf + ω µ) − ω Σωµ µ + (1 − γ) 0 2 = 0. (16) (Rf + ω µ) Let 0 −1 0 ω Σω J := µΣ µ,X := Rf + ω µ, and Y := . (17) 0 2 (Rf + ω µ) Multiplying (16) by ω0 and µ0Σ−1 we have

X − Rf + (1 − γ)Rf Y = 0,

X − R  J + (1 − γ) f − YJ = 0. X Next, substituting Y in the second equation for Y in the first one it holds:

X − R X − R  J + (1 − γ) f + f J = 0 X Rf (1 − γ) and, consequently,

JRf X + (1 − γ)Rf (X − Rf ) + (X − Rf )XJ = 0, (18)

13 or 2 2 JX + (1 − γ)Rf X − (1 − γ)Rf = 0, (19) The roots of the quadratic equation (19) with respect to X are given by √ γ − 1 ± D X = R , (20) ± f 2J where D = (γ − 1)2 − 4(γ − 1)J, and Rf − X± Y± = . (21) Rf (1 − γ) Moreover, equation (20) shows that the quadratic equation has a solution if and only if D ≥ 0 which is equivalent to γ ∈ (0, 1] ∪ [1 + 4J, ∞). Since the FOCs only provide the critical points, it is necessary to discover a maximum of the objective function (14). In order to do that, it is sufficient to compare the argument of the exponent in (14) for both combinations of (X+,Y+) and (X−,Y−):

(1 − γ)2 (1 − γ)2 (1 − γ) ln X + Y − (1 − γ) ln X − Y + 2 + − 2 −   (21) X+ X− − X+ = (1 − γ) ln + (22) X− 2Rf " √ √ # (20) γ − 1 + D D = (1 − γ) ln √ − γ − 1 − D 2J

Next we will see that (22) is positive for γ > 1 + 4J: " √ √ # ∂ γ − 1 + D D ln √ − ∂γ γ − 1 − D 2J  ∂  √  ∂  √ ∂γ D   ∂γ D   1 + √ γ − 1 − D − 1 − √ γ − 1 + D ∂ 2 D 2 D D (23) = − ∂γ√ 4(γ − 1)J 4J D ∂ D √ ∂ D √ 2 ∂γ√ (γ − 1) − 2 D ∂γ√ (γ − 1) −2 D = 2 D − D = . 4(γ − 1)J 4J(γ − 1) 4J(γ − 1)

Meanwhile " √ √ # γ − 1 + D D 4J ln √ − = ln = 0. (24) γ − 1 − D 2J 4J γ=1+4J

Thus, the second multiplier in (22) is a monotonically decreasing and a negative function of γ for γ > 1 + 4J and, thus, (22) is positive. Hence, fro all γ > 1 + 4J the maximum of (14) is attained at (X−,Y−). 0 It is notable that Rf + ω µ or X is required to be positive, but for 0 < γ < 1, X− becomes negative, consequently (X+,Y+) is an only candidate for an extrema point. In

14 order to investigate the type of the extrema we calculate the second derivative of the objective function given by ω and analyse a sign of the quadratic form:

2 " 0 # ∂ 0 1 − γ ω Σω ln (Rf + ω µ) + 2 0 2 ∂ω 2 (Rf + ω µ)  0 0  (16) ∂ Σω (Rf + ω µ) − ω Σωµ = µ + (1 − γ) 0 2 ∂ω (Rf + ω µ) 0 (21) ∂  Σω ω µ  = (1 − γ) 0 + µ ∂ω Rf + ω µ Rf (1 − γ) " # Σ (R + ω0µ) − Σωµ0 µµ0 = (1 − γ) f + 0 2 (Rf + ω µ) Rf (1 − γ) (25)   1  0  Y − Xµµ 0 (16) Σ 1−γ µµ = (1 − γ)  − 2 +  X X Rf (1 − γ)

  1 X 1  0  − − Xµµ 0 (21) Σ 1−γ Rf (1−γ) 1−γ µµ = (1 − γ)  − 2 +  X X Rf (1 − γ)  Σ µµ0  = (1 − γ) + 2 . X Rf (1 − γ)

Thus, the quadratic form is positive definite for γ ∈ (0, 1) and in this case (X+,Y+) provide a minima of the value function (14). As the result, given by (16), (17), (20) and (21): X2 ω∗ = Σ−1µ , (26) Rf (γ − 1) and the single period approximate optimal portfolio wights are:  2 γ − 1 − p(γ − 1)2 − 4(γ − 1)µ0Σ−1µ ∗ −1 ω = Σ µRf . (27) 4(γ − 1) µ0Σ−1µ2

Proof of the Corollary 2.

" p 2 # ∂ (17),(20) ∂ γ − 1 − (γ − 1) − 4(γ − 1)J ω∗0µ = R ∂γ ∂γ f 2J ! R 2(γ − 1) − 4J = f 1 − 2J 2p(γ − 1)2 − 4(γ − 1)J R p(γ − 1)2 − 4(γ − 1)J − (γ − 1) + 2J = f 2J p(γ − 1)2 − 4(γ − 1)J R (γ − 1)2 − 4(γ − 1)J − (γ − 1 − 2J)2 = √f 2J D p(γ − 1)2 − 4(γ − 1)J + γ − 1 − 2J R −4J 2 γ>1+4J = √f < 0. 2J D p(γ − 1)2 − 4(γ − 1)J + γ − 1 − 2J

15 Taking into account that ω∗0µ is positive definite for γ > 1 + 4J:

p 2 (17),(20) γ − 1 − (γ − 1) − 4(γ − 1)J ω∗0µ = R − R f 2J f γ − 1 − 2J − p(γ − 1)2 − 4(γ − 1)J = R f 2J R (γ − 1 − 2J)2 − (γ − 1)2 + 4(γ − 1)J = f 2J γ − 1 − 2J + p(γ − 1)2 − 4(γ − 1)J R 4J 2 γ>1+4J = f > 0, 2J γ − 1 − 2J + p(γ − 1)2 − 4(γ − 1)J we have that portfolio variance also decreases on γ as a square of ω∗0µ. Proof of the Corollary 3. Multiplying (16) by ω0 and 10Σ−1 and using (17) we have:

X − Rf + (1 − γ)Rf Y = 0,  1  10Σ−1µ + (1 − γ) − Y 10Σ−1µ = 0. X Next, substituting Y from the second equation for Y in the first one it holds:  1 X − R  1Σ−1µ + (1 − γ) + f 1Σ−1µ = 0 X Rf (1 − γ) and, consequently, (γ − 1)R X2 = f . 10Σ−1µ Thus, according to (26), the tangency portfolio has its weights being

Σ−1µ ω = . tgc 10Σ−1µ At the same time

 0 −1 2 0 0 2 (γtgc − 1)Rf µ Σ µ 1 Σµ (ωtgcµ + Rf ) = 0 −1 ⇒ γtgc = 0 −1 + Rf + 1. 1 Σ µ 1 Σ µ Rf

References

Barberis, N. (2000). Investing for the long run when returns are predictable. The Journal of Finance, 55(1):225–264. Bodnar, T., Parolya, N., and Schmid, W. (2015a). A closed-form solution of the multi- period portfolio choice problem for a quadratic utility function. Annals of Operations Research, 229(1):121–158. Bodnar, T., Parolya, N., and Schmid, W. (2015b). On the exact solution of the multi- period portfolio choice problem for an exponential utility under return predictability. European Journal of Operational Research, 246(2):528–542.

16 Brandt, M. (2009). Portfolio choice problems. Handbook of financial econometrics, 1:269– 336.

Brandt, M. W., Goyal, A., Santa-Clara, P., and Stroud, J. R. (2005). A simulation approach to dynamic portfolio choice with an application to learning about return predictability. The Review of Financial Studies, 18(3):831–873.

Brandt, M. W. and Santa-Clara, P. (2006). Dynamic portfolio selection by augmenting the asset space. The Journal of Finance, 61(5):2187–2217.

Broadie, M. and Shen, W. (2017). Numerical solutions to dynamic portfolio problems with upper bounds. Computational Management Science, 14(2):215–227.

Campbell, J. Y. and Viceira, L. M. (2002). Strategic asset allocation: portfolio choice for long-term investors. Clarendon Lectures in Economic.

C¸anako˘glu,E. and Ozekici,¨ S. (2010). Portfolio selection in stochastic markets with hara utility functions. European Journal of Operational Research, 201(2):520–536.

C¸anako˘glu,E. and Ozekici,¨ S. (2012). Hara frontiers of optimal portfolios in stochastic markets. European Journal of Operational Research, 221(1):129–137.

C¸etinkaya, E. and Thiele, A. (2016). A moment matching approach to log-normal portfolio optimization. Computational Management Science, 13(4):501–520.

Herzel, S. and Nicolosi, M. (2019). Optimal strategies with option compensation under mean reverting returns or volatilities. Computational Management Science, 16(1-2):47– 69.

Markowitz, H. (1952). Portfolio selection. The Journal of Finance, 7(1):77–91.

Merton, R. C. (1972). An analytic derivation of the efficient portfolio frontier. Journal of financial and quantitative analysis, pages 1851–1872.

Pennacchi, G. G. (2008). Theory of asset pricing. Pearson/Addison-Wesley Boston.

Taras, B., Dmytro, I., Nestor, P., and Schmid, W. (2020). Mean-variance efficiency of op- timal power and logarithmic utility portfolios. Mathematics and Financial , 14(4):675–698.

17