Portfolio optimization

Last updated

Portfolio optimization is the process of selecting an optimal portfolio (asset distribution), out of a set of considered portfolios, according to some objective. The objective typically maximizes factors such as expected return, and minimizes costs like financial risk, resulting in a multi-objective optimization problem. Factors being considered may range from tangible (such as assets, liabilities, earnings or other fundamentals) to intangible (such as selective divestment).

Contents

Modern portfolio theory

Modern portfolio theory was introduced in a 1952 doctoral thesis by Harry Markowitz, where the Markowitz model was first defined. [1] [2] The model assumes that an investor aims to maximize a portfolio's expected return contingent on a prescribed amount of risk. Portfolios that meet this criterion, i.e., maximize the expected return given a prescribed amount of risk, are known as efficient portfolios. By definition, any other portfolio yielding a higher amount of expected return must also have excessive risk. This results in a trade-off between the desired expected return and allowable risk. This risk-expected return relationship of efficient portfolios is graphically represented by a curve known as the efficient frontier. All efficient portfolios, each represented by a point on the efficient frontier, are well-diversified. While ignoring higher moments of the return can lead to significant over-investment in risky securities, especially when volatility is high, [3] the optimization of portfolios when return distributions are non-Gaussian is mathematically challenging. [4]

Optimization methods

The portfolio optimization problem is specified as a constrained utility-maximization problem. Common formulations of portfolio utility functions define it as the expected portfolio return (net of transaction and financing costs) minus a cost of risk. The latter component, the cost of risk, is defined as the portfolio risk multiplied by a risk aversion parameter (or unit price of risk). For return distributions that are Gaussian, this is equivalent to maximizing a certain quantile of the return, where the corresponding probability is dictated by the risk aversion parameter. Practitioners often add additional constraints to improve diversification and further limit risk. Examples of such constraints are asset, sector, and region portfolio weight limits.

Specific approaches

Portfolio optimization often takes place in two stages: optimizing weights of asset classes to hold, and optimizing weights of assets within the same asset class. An example of the former would be choosing the proportions placed in equities versus bonds, while an example of the latter would be choosing the proportions of the stock sub-portfolio placed in stocks X, Y, and Z. Equities and bonds have fundamentally different financial characteristics and have different systematic risk and hence can be viewed as separate asset classes; holding some of the portfolio in each class provides some diversification, and holding various specific assets within each class affords further diversification. By using such a two-step procedure one eliminates non-systematic risks both on the individual asset and the asset class level. For the specific formulas for efficient portfolios, [5] see Portfolio separation in mean-variance analysis.

One approach to portfolio optimization is to specify a von Neumann–Morgenstern utility function defined over final portfolio wealth; the expected value of utility is to be maximized. To reflect a preference for higher rather than lower returns, this objective function is increasing in wealth, and to reflect risk aversion it is concave. For realistic utility functions in the presence of many assets that can be held, this approach, while theoretically the most defensible, can be computationally intensive.

Harry Markowitz [6] developed the "critical line method", a general procedure for quadratic programming that can handle additional linear constraints and upper and lower bounds on holdings. Moreover, in this context, the approach provides a method for determining the entire set of efficient portfolios. Its application here was later explicated by William Sharpe. [7]

Mathematical tools

The complexity and scale of optimizing portfolios over many assets means that the work is generally done by computer. Central to this optimization is the construction of the covariance matrix for the rates of return on the assets in the portfolio.

Techniques include:

Optimization constraints

Portfolio optimization is usually done subject to constraints, such as regulatory constraints, or illiquidity. These constraints can lead to portfolio weights that focus on a small sub-sample of assets within the portfolio. When the portfolio optimization process is subject to other constraints such as taxes, transaction costs, and management fees, the optimization process may result in an under-diversified portfolio. [14]

Regulation and taxes

Investors may be forbidden by law to hold some assets. In some cases, unconstrained portfolio optimization would lead to short-selling of some assets. However short-selling can be forbidden. Sometimes it is impractical to hold an asset because the associated tax cost is too high. In such cases appropriate constraints must be imposed on the optimization process.

Transaction costs

Transaction costs are the costs of trading to change the portfolio weights. Since the optimal portfolio changes with time, there is an incentive to re-optimize frequently. However, too frequent trading would incur too-frequent transactions costs; so the optimal strategy is to find the frequency of re-optimization and trading that appropriately trades off the avoidance of transaction costs with the avoidance of sticking with an out-of-date set of portfolio proportions. This is related to the topic of tracking error, by which stock proportions deviate over time from some benchmark in the absence of re-balancing.

Concentration risk

Concentration risk refers to the risk caused by holding an exposure to a single position or sector that is large enough to cause material losses to the overall portfolio when adverse events occur. If the portfolio is optimized without any constraints with regards to concentration risk, the optimal portfolio can be any risky-asset portfolio, and therefore there is nothing to prevent it from being a portfolio that invests solely in a single asset. Managing concentration risk should be part of a comprehensive risk management framework [15] and to achieve a reduction in such a risk it is possible to add constraints that force upper bound limits to the weight that can be attributed to any single component of the optimal portfolio.

Improving portfolio optimization

Correlations and risk evaluation

Different approaches to portfolio optimization measure risk differently. In addition to the traditional measure, standard deviation, or its square (variance), which are not robust risk measures, other measures include the Sortino ratio, CVaR (Conditional Value at Risk), and statistical dispersion.

Investment is a forward-looking activity, and thus the covariances of returns must be forecast rather than observed.

Portfolio optimization assumes the investor may have some risk aversion and the stock prices may exhibit significant differences between their historical or forecast values and what is experienced. In particular, financial crises are characterized by a significant increase in correlation of stock price movements which may seriously degrade the benefits of diversification. [16]

In a mean-variance optimization framework, accurate estimation of the variance-covariance matrix is paramount. Quantitative techniques that use Monte-Carlo simulation with the Gaussian copula and well-specified marginal distributions are effective. [17] Allowing the modeling process to allow for empirical characteristics in stock returns such as autoregression, asymmetric volatility, skewness, and kurtosis is important. Not accounting for these attributes can lead to severe estimation error in the correlations, variances and covariances that have negative biases (as much as 70% of the true values). [18]

Other optimization strategies that focus on minimizing tail-risk (e.g., value at risk, conditional value at risk) in investment portfolios are popular among risk averse investors. To minimize exposure to tail risk, forecasts of asset returns using Monte-Carlo simulation with vine copulas to allow for lower (left) tail dependence (e.g., Clayton, Rotated Gumbel) across large portfolios of assets are most suitable. [19] (Tail) risk parity focuses on allocation of risk, rather than allocation of capital.

More recently, hedge fund managers have been applying "full-scale optimization" whereby any investor utility function can be used to optimize a portfolio. [20] It is purported that such a methodology is more practical and suitable for modern investors whose risk preferences involve reducing tail risk, minimizing negative skewness and fat tails in the returns distribution of the investment portfolio. [21] Where such methodologies involve the use of higher-moment utility functions, it is necessary to use a methodology that allows for forecasting of a joint distribution that accounts for asymmetric dependence. A suitable methodology that allows for the joint distribution to incorporate asymmetric dependence is the Clayton Canonical Vine Copula. See Copula (probability theory) § Quantitative finance.

Cooperation in portfolio optimization

A group of investors, instead of investing individually, may choose to invest their total capital into the joint portfolio, and then divide the (uncertain) investment profit in a way which suits best their utility/risk preferences. It turns out that, at least in the expected utility model, [22] and mean-deviation model, [23] each investor can usually get a share which he/she values strictly more than his/her optimal portfolio from the individual investment.

See also

Related Research Articles

Financial economics is the branch of economics characterized by a "concentration on monetary activities", in which "money of one type or another is likely to appear on both sides of a trade". Its concern is thus the interrelation of financial variables, such as share prices, interest rates and exchange rates, as opposed to those concerning the real economy. It has two main areas of focus: asset pricing and corporate finance; the first being the perspective of providers of capital, i.e. investors, and the second of users of capital. It thus provides the theoretical underpinning for much of finance.

<span class="mw-page-title-main">Capital asset pricing model</span> Model used in finance

In finance, the capital asset pricing model (CAPM) is a model used to determine a theoretically appropriate required rate of return of an asset, to make decisions about adding assets to a well-diversified portfolio.

Market risk is the risk of losses in positions arising from movements in market variables like prices and volatility. There is no unique classification as each classification may refer to different aspects of market risk. Nevertheless, the most commonly used types of market risk are:

In economics and finance, risk neutral preferences are preferences that are neither risk averse nor risk seeking. A risk neutral party's decisions are not affected by the degree of uncertainty in a set of outcomes, so a risk neutral party is indifferent between choices with equal expected payoffs even if one choice is riskier.

Harry Max Markowitz was an American economist who received the 1989 John von Neumann Theory Prize and the 1990 Nobel Memorial Prize in Economic Sciences.

Modern portfolio theory (MPT), or mean-variance analysis, is a mathematical framework for assembling a portfolio of assets such that the expected return is maximized for a given level of risk. It is a formalization and extension of diversification in investing, the idea that owning different kinds of financial assets is less risky than owning only one type. Its key insight is that an asset's risk and return should not be assessed by itself, but by how it contributes to a portfolio's overall risk and return. The variance of return is used as a measure of risk, because it is tractable when assets are combined into portfolios. Often, the historical variance and covariance of returns is used as a proxy for the forward-looking versions of these quantities, but other, more sophisticated methods are available.

In finance, a portfolio is a collection of investments.

In finance, diversification is the process of allocating capital in a way that reduces the exposure to any one particular asset or risk. A common path towards diversification is to reduce risk or volatility by investing in a variety of assets. If asset prices do not change in perfect synchrony, a diversified portfolio will have less variance than the weighted average variance of its constituent assets, and often less volatility than the least volatile of its constituents.

The following outline is provided as an overview of and topical guide to finance:

Simply stated, post-modern portfolio theory (PMPT) is an extension of the traditional modern portfolio theory (MPT) of Markowitz and Sharpe. Both theories provide analytical methods for rational investors to use diversification to optimize their investment portfolios. The essential difference between PMPT and MPT is that PMPT emphasizes the return that must be earned on an investment in order to meet future, specified obligations, MPT is concened only with the absolute return vis-a-vis the risk-free rate.

In investment portfolio construction, an investor or analyst is faced with determining which asset classes, such as domestic fixed income, domestic equity, foreign fixed income, and foreign equity, to invest in and what proportion of the total portfolio should be of each asset class. Harry Markowitz (1959) first described a method for constructing a portfolio with optimal risk/return characteristics. His portfolio optimization method finds the minimum risk portfolio with a given expected return. Because the Markowitz or Mean-Variance Efficient Portfolio is calculated from the sample mean and covariance, which are likely different from the population mean and covariance, the resulting investment portfolio may allocate too much weight to assets with better estimated than true risk/return characteristics. To account for the uncertainty of the sample estimates, a financial analyst can create many alternative efficient frontiers based on resampled versions of the data. Each resampled dataset will result in a different set of Markowitz efficient portfolios. These efficient frontiers of portfolios can then be averaged to create a resampled efficient frontier. The appropriate compromise between the investor's Risk aversion and desired return will then guide the financial analyst to choose a portfolio from the set of resampled efficient frontier portfolios. Since such a portfolio is different from the Markowitz efficient portfolio it will have suboptimal risk/return characteristics with respect to the sample mean and covariance, but optimal characteristics when averaged over the many possible values of the unknown true mean and covariance. Resampled Efficiency is covered by U. S. patent #6,003,018, patent pending worldwide. New Frontier Advisors, LLC, has exclusive worldwide licensing rights.

In finance, the Black–Litterman model is a mathematical model for portfolio allocation developed in 1990 at Goldman Sachs by Fischer Black and Robert Litterman, and published in 1992. It seeks to overcome problems that institutional investors have encountered in applying modern portfolio theory in practice. The model starts with an asset allocation based on the equilibrium assumption and then modifies that allocation by taking into account the opinion of the investor regarding future asset performance.

Goals-Based Investing or Goal-Driven Investing is the use of financial markets to fund goals within a specified period of time. Traditional portfolio construction balances expected portfolio variance with return and uses a risk aversion metric to select the optimal mix of investments. By contrast, GBI optimizes an investment mix to minimize the probability of failing to achieve a minimum wealth level within a set period of time.

<span class="mw-page-title-main">Efficient frontier</span> Investment portfolio which occupies the "efficient" parts of the risk-return spectrum

In modern portfolio theory, the efficient frontier is an investment portfolio which occupies the "efficient" parts of the risk–return spectrum. Formally, it is the set of portfolios which satisfy the condition that no other portfolio exists with a higher expected return but with the same standard deviation of return. The efficient frontier was first formulated by Harry Markowitz in 1952; see Markowitz model.

In portfolio theory, a mutual fund separation theorem, mutual fund theorem, or separation theorem is a theorem stating that, under certain conditions, any investor's optimal portfolio can be constructed by holding each of certain mutual funds in appropriate ratios, where the number of mutual funds is smaller than the number of individual assets in the portfolio. Here a mutual fund refers to any specified benchmark portfolio of the available assets. There are two advantages of having a mutual fund theorem. First, if the relevant conditions are met, it may be easier for an investor to purchase a smaller number of mutual funds than to purchase a larger number of assets individually. Second, from a theoretical and empirical standpoint, if it can be assumed that the relevant conditions are indeed satisfied, then implications for the functioning of asset markets can be derived and tested.

In finance, marginal conditional stochastic dominance is a condition under which a portfolio can be improved in the eyes of all risk-averse investors by incrementally moving funds out of one asset and into another. Each risk-averse investor is assumed to maximize the expected value of an increasing, concave von Neumann-Morgenstern utility function. All such investors prefer portfolio B over portfolio A if the portfolio return of B is second-order stochastically dominant over that of A; roughly speaking this means that the density function of A's return can be formed from that of B's return by pushing some of the probability mass of B's return to the left and then spreading out some of the density mass.

Mathematical finance, also known as quantitative finance and financial mathematics, is a field of applied mathematics, concerned with mathematical modeling of financial markets.

Intertemporal portfolio choice is the process of allocating one's investable wealth to various assets, especially financial assets, repeatedly over time, in such a way as to optimize some criterion. The set of asset proportions at any time defines a portfolio. Since the returns on almost all assets are not fully predictable, the criterion has to take financial risk into account. Typically the criterion is the expected value of some concave function of the value of the portfolio after a certain number of time periods—that is, the expected utility of final wealth. Alternatively, it may be a function of the various levels of goods and services consumption that are attained by withdrawing some funds from the portfolio after each time period.

Philippe J.S. De Brouwer is a European investment and banking professional as well as academician in finance and investing. As a scientist he is mostly known for his solution to the Fallacy of Large Numbers and his formulation of the Maslowian Portfolio Theory in the field of investment advice.

In finance, the Markowitz model ─ put forward by Harry Markowitz in 1952 ─ is a portfolio optimization model; it assists in the selection of the most efficient portfolio by analyzing various possible portfolios of the given securities. Here, by choosing securities that do not 'move' exactly together, the HM model shows investors how to reduce their risk. The HM model is also called mean-variance model due to the fact that it is based on expected returns (mean) and the standard deviation (variance) of the various portfolios. It is foundational to Modern portfolio theory.

References

  1. Markowitz, H.M. (March 1952). "Portfolio Selection". The Journal of Finance. 7 (1): 77–91. doi:10.2307/2975974. JSTOR   2975974.
  2. Markowitz, H.M. (1959). Portfolio Selection: Efficient Diversification of Investments. New York: John Wiley & Sons. (reprinted by Yale University Press, 1970, ISBN   978-0-300-01372-6; 2nd ed. Basil Blackwell, 1991, ISBN   978-1-55786-108-5)
  3. Cvitanić, Jakša; Polimenis, Vassilis; Zapatero, Fernando (1 January 2008). "Optimal portfolio allocation with higher moments". Annals of Finance. 4 (1): 1–28. doi:10.1007/s10436-007-0071-5. ISSN   1614-2446. S2CID   16514619.
  4. Kim, Young Shin; Giacometti, Rosella; Rachev, Svetlozar; Fabozzi, Frank J.; Mignacca, Domenico (21 November 2012). "Measuring financial risk and portfolio optimization with a non-Gaussian multivariate model". Annals of Operations Research. 201 (1): 325–343. doi:10.1007/s10479-012-1229-8. S2CID   45585936.
  5. Merton, Robert. September 1972. "An analytic derivation of the efficient portfolio frontier," Journal of Financial and Quantitative Analysis 7, 1851–1872.
  6. Markowitz, Harry (1956). "The optimization of a quadratic function subject to linear constraints". Naval Research Logistics Quarterly . 3 (1–2): 111–133. doi:10.1002/nav.3800030110.
  7. The Critical Line Method in William Sharpe, Macro-Investment Analysis (online text)
  8. Rockafellar, R. Tyrrell; Uryasev, Stanislav (2000). "Optimization of conditional value-at-risk" (PDF). Journal of Risk. 2 (3): 21–42. doi:10.21314/JOR.2000.038. S2CID   854622.
  9. Kapsos, Michalis; Zymler, Steve; Christofides, Nicos; Rustem, Berç (Summer 2014). "Optimizing the Omega Ratio using Linear Programming" (PDF). Journal of Computational Finance. 17 (4): 49–57. doi:10.21314/JCF.2014.283.
  10. Talebi, Arash; Molaei, Sheikh (17 September 2010). "Performance investigation and comparison of two evolutionary algorithms in portfolio optimization: Genetic and particle swarm optimization". 2010 2nd IEEE International Conference on Information and Financial Engineering. pp. 430–437. doi:10.1109/icife.2010.5609394. ISBN   978-1-4244-6927-7. S2CID   17386345.
  11. Shapiro, Alexander; Dentcheva, Darinka; Ruszczyński, Andrzej (2009). Lectures on stochastic programming: Modeling and theory (PDF). MPS/SIAM Series on Optimization. Vol. 9. Philadelphia, PA: Society for Industrial and Applied Mathematics (SIAM). Mathematical Programming Society (MPS). pp. xvi+436. ISBN   978-0-89871-687-0. MR   2562798.
  12. Zhu, Zhe; Welsch, Roy E. (2018). "Robust dependence modeling for high-dimensional covariance matrices with financial applications". Ann. Appl. Stat. 12 (2): 1228–1249. doi: 10.1214/17-AOAS1087 . S2CID   23490041.
  13. Sefiane, Slimane and Benbouziane, Mohamed (2012). Portfolio Selection Using Genetic Algorithm Archived 29 April 2016 at the Wayback Machine , Journal of Applied Finance & Banking, Vol. 2, No. 4 (2012): pp. 143–154.
  14. Humphrey, J.; Benson, K.; Low, R.K.Y.; Lee, W.L. (2015). "Is diversification always optimal?" (PDF). Pacific Basin Finance Journal. 35 (B): B. doi:10.1016/j.pacfin.2015.09.003.
  15. "Concentrate on Concentration Risk | FINRA.org". www.finra.org. 15 June 2022. Retrieved 16 March 2024.
  16. Chua, D.; Krizman, M.; Page, S. (2009). "The Myth of Diversification". Journal of Portfolio Management. 36 (1): 26–35. doi:10.3905/JPM.2009.36.1.026. S2CID   154921810.
  17. Low, R.K.Y.; Faff, R.; Aas, K. (2016). "Enhancing mean–variance portfolio selection by modeling distributional asymmetries" (PDF). Journal of Economics and Business. 85: 49–72. doi:10.1016/j.jeconbus.2016.01.003.
  18. Fantazzinni, D. (2009). "The effects of misspecified marginals and copulas on computing the value at risk: A Monte Carlo study". Computational Statistics & Data Analysis. 53 (6): 2168–2188. doi:10.1016/j.csda.2008.02.002.
  19. Low, R.K.Y.; Alcock, J.; Faff, R.; Brailsford, T. (2013). "Canonical vine copulas in the context of modern portfolio management: Are they worth it?" (PDF). Journal of Banking & Finance. 37 (8): 3085. doi:10.1016/j.jbankfin.2013.02.036. S2CID   154138333.
  20. Chua, David; Kritzman, Mark; Page, Sebastien (2009). "The Myth of Diversification". Journal of Portfolio Management. 36 (1): 26–35. doi:10.3905/JPM.2009.36.1.026. S2CID   154921810.
  21. Adler, Tim; Kritzman, Mark (2007). "Mean-Variance versus Full-Scale Optimization: In and Out of Sample". Journal of Asset Management. 7 (5): 71–73. doi:10.2469/dig.v37.n3.4799.
  22. Xia, Jianming (2004). "Multi-agent investment in incomplete markets". Finance and Stochastics. 8 (2): 241–259. doi:10.1007/s00780-003-0115-2. S2CID   7162635.
  23. Grechuk, B., Molyboha, A., Zabarankin, M. (2013). "Cooperative games with general deviation measures", Mathematical Finance, 23(2), 339–365.

Bibliography