Skip to main content
Log in

Tail risk in pension funds: an analysis using ARCH models and bilinear processes

  • Original Research
  • Published:
Review of Quantitative Finance and Accounting Aims and scope Submit manuscript

Abstract

Pension funding rules and practice contain implicit smoothing and counter-cyclical mechanisms. We set up a stylized model to investigate whether this may give rise to tail risk, in the form of large but rare losses, when pension liabilities are imperfectly but optimally hedged by pension fund assets. We find that pension losses follow a nonlinear dynamic process, and we derive a complete description of the stochastic properties of this process using Markov chain and bilinear stochastic process theory. The resulting pension dynamics resembles that of a modified ARCH model, which suggests that bursts in volatility may occur, and tail risk may be present. Simulations confirm that pension losses exhibit skewness, leptokurtosis and heavy tails, specially when cash flow smoothing is pronounced. Regulators and investors should be aware of the total amount of smoothing in pension funds as this may contribute to extreme losses, which may adversely affect the security of employee benefits as well as the valuation of firms with corporate pension plans.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. Global pension funds experienced losses of about 25 % of their asset value, as a percentage of pension liability, in 2008 (Towers Watson 2011).

  2. Data from Towers Watson (2011). In ‘defined benefit’ pension plans, employees receive pensions at retirement that are a function of their service and salary, with the employer taking on the investment and longevity risk associated with these benefits. This is in contrast with ‘defined contribution’ plans where employees take on all the risks. Contributions from firms and employees to corporate defined benefit funds amounted to about 0.5 % of US GDP in 2009 (OECD data reported in Yermo and Severinson, 2010). Our analysis is focused on corporate pension plans, but may be extended to plans covering government employees. In the US in 2008, state government defined benefit plans had assets of $2.3 trillion (Novy-Marx and Rauh 2011).

  3. This is typically based on the yields on high-quality corporate bonds to reflect the credit risk present in pension benefits. Practical examples of such yield curves, updated monthly, are supplied by Citigroup (2010) and Mercer (2012). The US Treasury Department also implements a high-quality corporate bond yield curve for use with the pension funding rules in the US Pension Protection Act 2006 (Girola 2011).

  4. The present value is calculated by discounting each future cash flow of $1 using the suitable spot rate. Allowance is made for mortality through a survival probability, since the annuity payout is discontinued at death. The term α(tx) in Eq. (1) is often referred to as an actuarial present value (Dickson et al. 2009, p. 76).

  5. See for example Sundaresan and Zapatero (1997), McGill et al. (2004, p. 235), Blake (2006b, p. 194) and Dickson et al. (2009, p. 306–312) for more details.

  6. The PBO is also the preferred measure of pension liability under several international accounting standards. An alternative measure of pension liability is the Accumulated Benefit Obligation (ABO) which is defined in FAS 87. It is also used by pension funding regulators in various countries because it is a termination measure, as compared with the PBO which is a broader, going-concern measure (Broeders and Chen 2010). Our subsequent analysis holds for the ABO rather than the PBO, as long as m x is re-defined as a function of the current salary of an individual, rather than his projected salary at retirement.

  7. This is a stylized version of actual pension cash flows. For a comprehensive and detailed description, from both funding and accounting perspectives, see chapters 22–26 of McGill et al. (2004).

  8. The Pensions Act 2004 in the United Kingdom requires explicit consideration of the strength of the sponsor covenant for regulatory funding purposes (Yermo and Severinson 2010; OECD 2007, p. 205).

  9. In theory, this may be achieved using a quadratic (variance-minimizing) hedge or an entropy-minimizing hedge (Duffie and Richardson 1991; Rouge and El Karaoui 2000). In practice, this is consistent with asset-liability management strategies, including Liability-Driven Investment (LDI) (Blake 2006b, p. 260–274). These are growing in popularity with pension plans worldwide, specially in the UK and in the Netherlands (OECD 2007, p. 61). Love et al. (2011) discuss the asset-liability mismatch prevailing in most US corporate pension plans but report that 51 % of them were instigating LDI programs in 2009.

  10. See the discussion of voluntary contributions in Sect. 2.2 and references therein. A detailed explanation of supplementary contributions, also known as ‘supplemental costs’, is given by McGill et al. (2004, p. 634).

  11. Note that this is a generalized setting which encapsulates amortization over p years if π j (for \(j=1, \ldots, p\)) is set equal to the reciprocal of the present value of a p-year annuity-due. In this case, we envisage that the dimension p of the loss vector equals the maximum recovery period (for example, p = 7 years under PPA 2006). However, p might be greater than the typical amortization period if funding practice also comprises smoothed liability and asset values.

  12. If log X ∼ N(μ, σ2), then \({\rm Var}{X} = e^{2\mu+\sigma^2} \left(e^{\sigma^2}-1\right)\). Substituting μ = 6.83 % and σ = 8.91 % gives a standard deviation of 9.6 %.

  13. Equivalently, the frontiers in Fig. 1 give the maximum m for various n. Note that \({m,n\in{\mathbb{N}}}\), so a polynomial curve is fitted to non-integer values for readability. Similar depictions appear elsewhere: see Bollerslev (1986, Fig. 1) in the econometrics literature on GARCH models, and Tong (1990, Fig. 4.10, p. 171) in the non-linear time series literature on bilinear processes.

  14. For a continuously distributed random variable X with cdf F X (x), the 95th percentile is F −1 X (0.95) and the TCE is \({{\mathbb{E}} \left[ {X} \; | \; {X>F_X^{-1}(0.95)} \right]}\). The TCE is closely related to other risk measures such as the Tail Value-at-Risk (TVaR), the conditional Value-at-Risk (CVaR) and the Expected Shortfall (ES). There are subtle distinctions but they are of no practical import here. For details, see Dowd (2002, p. 32) and references therein.

References

  • Adams J, Smith DJ (2009) Mind the gap: using derivatives overlays to hedge pension duration. Financ Anal J 65(4):60–67

    Article  Google Scholar 

  • ASB (2007a) Actuarial standard of practice no. 4. Measuring pension obligations and determining pension plan costs or contributions. Actuarial Standards Board, Washington

    Google Scholar 

  • ASB (2007b) Actuarial standard of practice no. 27. Selection of economic assumptions for measuring pension obligations. Actuarial Standards Board, Washington

    Google Scholar 

  • ASB (2009) Actuarial standard of practice no. 44. Selection and use of asset valuation methods for pension valuations. Actuarial Standards Board, Washington

    Google Scholar 

  • Bergstresser D, Desai MA, Rauh JD (2006) Earnings manipulation, pension assumptions, and managerial investment decisions. Q J Econ 121:157–195

    Article  Google Scholar 

  • Berkelaar A, Kouwenberg R (2003) Retirement saving with contribution payments and labor income as a benchmark for investments. J Econ Dyn Control 27:1069–1097

    Article  Google Scholar 

  • Blake D (2006a) Pension Economics. Wiley, Chichester

    Google Scholar 

  • Blake D (2006b) Pension Finance. John Wiley, Chichester

    Google Scholar 

  • Bollerslev T (1986) Generalized autoregressive conditional heteroscedasticity. J Econom 31:307–327

    Article  Google Scholar 

  • Bougerol P, Picard N (1992a) Strict stationarity of generalized autoregressive processes. Ann Probab 20(4):1714–1730

    Article  Google Scholar 

  • Bougerol P, Picard N (1992b) Stationarity of GARCH processes and of some nonnegative time series. J Econom 52:115–127

    Article  Google Scholar 

  • Broeders D, Chen A (2010) Pension regulation and the market value of pension liabilities: a contingent claims analysis using Parisian options. J Bank Financ 34(6):1201–1214

    Article  Google Scholar 

  • Campbell JL, Dhaliwal DS, Schwartz WC Jr (2010) Equity valuation effects of the Pension Protection Act of 2006. Contemp Acc Res 27(2):469–536

    Article  Google Scholar 

  • Campbell JL, Dhaliwal DS, Schwartz WC Jr (2012) Financing constraints and the cost of capital: evidence from the funding of corporate pension plans. Rev Financ Stud 25(3):868–912

    Article  Google Scholar 

  • Citigroup (2010) Citigroup pension liability index—revised methodology. http://www.soa.org/Files/Sections/pen-cpli-methodology.pdf. Accessed 10 July 2012

  • Committee on Retirement Systems Research (2001) Survey of asset valuation methods for defined benefit pension plans. Pension Forum 13(1):1–49. Society of Actuaries, Schaumburg

    Google Scholar 

  • Davidson J (1994) Stochastic limit theory. Oxford University Press, Oxford

    Google Scholar 

  • Dickson DCM, Hardy MR, Waters HR (2009) Actuarial mathematics for life contingent risks. Cambridge University Press, Cambridge

    Google Scholar 

  • Dowd K (2002) Measuring market risk. Wiley, Chichester

    Google Scholar 

  • Duffie D, Richardson HR (1991) Mean-variance hedging in continuous time. Ann Appl Probab 1:1–15

    Article  Google Scholar 

  • EIOPA (2012) Solvency II. European Insurance and Occupational Pensions Authority (EIOPA), Frankfurt am Main. https://eiopa.europa.eu/activities/insurance/solvency-ii/. Accessed 12 July 2012

  • Fan J, Yao Q (2003) Nonlinear time series: nonparametric and parametric methods. Springer, Berlin

    Google Scholar 

  • FASB (1985) Statement of financial accounting standards no. 87: employers’ accounting for pensions. Financial Accounting Standards Board (FASB), Stamford

  • FASB (2006) Statement of financial accounting standards no. 158: employers’ accounting for defined benefit pensions and other postretirement plans. Financial Accounting Standards Board (FASB), Stamford

  • Feller W (1971) An introduction to probability theory and its applications, vol. 2, 2nd ed. Wiley, London

    Google Scholar 

  • Franzoni F (2009) Underinvestment vs. overinvestment: evidence from price reactions to pension contributions. J Financ Econ 92:491–518

    Article  Google Scholar 

  • Franzoni F, Marín JM (2006) Pension plan funding and stock market efficiency. J Financ 61(2):921–956

    Article  Google Scholar 

  • Giraitis L, Leipus R, Robinson PM, Surgailis D (2004) LARCH, leverage and long memory. J Financ Econom 2(2):177–210

    Google Scholar 

  • Giraitis L, Robinson PM, Surgailis D (2000) A model for long memory conditional heteroscedasticity. Ann Appl Probab 10(3):1002–1024

    Google Scholar 

  • Girola JA (2011) The HQM yield curve: basic concepts. US Department of the Treasury. http://www.treasury.gov/resource-center/economic-policy/corp-bond-yield/Documents/ycp_oct2011.pdf. Accessed 10 July 2012

  • Grant CT, Grant GH, Ortega WR (2007) FASB’s quick fix for pension accounting is only first step. Financ Anal J 63:21–35

    Google Scholar 

  • Hann RN, Heflin F, Subramanayam KR (2007) Fair-value pension accounting. J Account Econ 44(3):328–358

    Google Scholar 

  • Horn RA, Johnson CR (1985) Matrix analysis. Cambridge University Press, Cambridge

    Google Scholar 

  • Jiang X (2011) The smoothing of pension expenses: a panel analysis. Rev Quant Financ Acc 37:451–476

    Google Scholar 

  • Josa-Fombellida R, Rincón-Zapatero JP (2006) Optimal investment decisions with a liability: the case of defined benefit pension plans. Insur Math Econ 39:81–98

    Google Scholar 

  • Kesten H (1973) Random difference equations and renewal theory for products of random matrices. Acta Math 131:207–248

    Google Scholar 

  • Kristensen D (2009) On stationarity and ergodicity of the bilinear model with applications to GARCH models. J Time Ser Anal 30(1):125–144

    Google Scholar 

  • Leibowitz ML, Henrikssan RD (1988) Portfolio optimization within a surplus framework. Financ Anal J 44(2):43–47

    Google Scholar 

  • Love DA, Smith PA, Wilcox DW (2011) The effect of regulation on optimal corporate pension risk. J Financ Econ 101:18–35

    Google Scholar 

  • McGill DM, Brown KN, Haley JJ, Schieber SJ (2004) Fundamentals of private pensions, 8th ed. Oxford University Press, Oxford

    Google Scholar 

  • Mercer (2012) Mercer pension discount yield curve and index rates in the US. http://www.mercer.com/articles/1213490. Accessed on 10 July 2012

  • Mitra S, Hossain M (2009) Value-relevance of pension transition adjustments and other comprehensive income components in the adoption year of SFAS No. 158. Rev Quant Financ Acc 33:279–301

    Google Scholar 

  • Novy-Marx R, Rauh J (2011) Public pension promises: how big are they and what are they worth? J Financ 66(4):1211–1249

    Google Scholar 

  • OECD (2007) Protecting pensions. Policy analysis and examples from OECD countries. Private pensions series no. 8, Organization for Economic Cooperation and Development. OECD Publishing, Paris

  • Owadally MI, Haberman S (1999) Pension fund dynamics and gains/losses due to random rates of investment return. N Am Actuar J 3(3):105–117

    Google Scholar 

  • Owadally MI, Haberman S (2004a) Efficient gain and loss amortization and optimal funding in pension plans. N Amer Actuar J 8(1):21–36

    Google Scholar 

  • Owadally MI, Haberman S (2004b) The treatment of assets in pension funding. ASTIN Bull 34(2):425–433

    Google Scholar 

  • Petersen MA (1994) Cash flow variability and firm’s pension choice. A role for operating leverage. J Financ Econ 36:361–383

    Google Scholar 

  • Pham DT (1985) Bilinear Markovian representation and bilinear models. Stoch Proc Appl 20:295–306

    Google Scholar 

  • Pham DT (1986) The mixing property of bilinear and generalised random coefficient autoregressive models. Stoch Proc Appl 23:291–300

    Google Scholar 

  • Pham DT (1993) Bilinear time series models. In: Tong H (ed) Dimension estimation and models. World Scientific, Singapore, pp. 191–223

    Google Scholar 

  • Picconi M (2006) The perils of pensions: does pension accounting lead investors and analysts astray? Acc Rev 81:925–955

    Google Scholar 

  • Rauh JD (2006) Investment and financing constraints: evidence from the funding of corporate pension plans. J Financ 61:33–71

    Google Scholar 

  • Rauh JD (2009) Risk shifting versus risk management: investment policy in corporate pension plans. Rev Financ Stud 22(7):2687–2733

    Google Scholar 

  • Rouge R, El Karoui N (2000) Pricing via utility maximization and entropy. Math Financ 10:259–276

    Google Scholar 

  • Rudolf M, Ziemba WT (2004) Intertemporal surplus management. J Econ Dyn Control 28(5):975–990

    Google Scholar 

  • Sharpe WF, Tint LG (1990) Liabilities—a new approach. J Portfolio Manage 16:5–10

    Google Scholar 

  • Shivdasani A, Stefanescu I (2010) How do pensions affect corporate capital structure decisions? Rev Financ Stud 23(3):1287–1323

    Google Scholar 

  • Sundaresan S, Zapatero F (1997) Valuation, optimal asset allocation and retirement incentives of pension plans. Rev Financ Stud 10:631–660

    Google Scholar 

  • Tapia W (2008) Comparing aggregate investment returns in privately managed pension funds: an initial assessment. OECD working papers on insurance and private pensions no. 21, Organization for Economic Cooperation and Development. OECD Publishing, Paris

  • Tong H (1990) Non-linear time series. A dynamical system approach. Clarendon Press, Oxford

    Google Scholar 

  • Towers Watson (2011) Global pension asset study. Towers Watson, New York

    Google Scholar 

  • Yermo J, Severinson C (2010) The impact of the financial crisis on defined benefit plans and the need for counter-cyclical funding regulations. OECD working papers on finance, insurance and private pensions no. 3, Organization for Economic Cooperation and Development. OECD Publishing, Paris

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Iqbal Owadally.

Appendices

Appendix: 1 Proof of approximation in Eq. (5)

Recall that Y is the liability when it is calculated at the term-independent discount rate δ. From Eq. (2),

$$ Y = \sum_{x=E}^R m_x \ell_x \widetilde{\alpha}(x), $$
(27)

where \(\widetilde{\alpha}(x)\) is the present value, at rate δ and at time t, of a deferred annuity paying $1 yearly in advance from retirement age R till death, for an individual aged x ≤ R at time t:

$$ \widetilde{\alpha}(x) = \sum_{j=0}^{\infty} e^{-(R-x+j)\delta} \ell_{R+j}/\ell_x. $$
(28)

The normal cash flow Z t is Z t  = Z (1) t  − Z (2) t  + Z (3) t , where each component is defined in Sect. 3.2.

$$ Z_t = m_R \ell_R \alpha(t,R) - \sum_{x=E}^{R-1} (m_{x+1}- m_x) \ell_x \alpha(t,x) + e^{-\delta} (\delta_t - \delta)Y_t. $$
(29)

By analogy to Y in Eq. (27), we may define Z to be the normal cash flow Z t when it is calculated using the term-independent discount rate δ.

$$ Z = m_R \ell_R \widetilde{\alpha}(R) - \sum_{x=E}^{R-1} (m_{x+1}- m_x) \ell_x \widetilde{\alpha}(x). $$
(30)

Noting that m E  = 0 by definition (no benefit right endowed at entry into the plan), the above may be rewritten as

$$ Z = \sum_{x=E}^{R} m_x \left[ \ell_x \widetilde{\alpha}(x) - \ell_{x-1} \widetilde{\alpha}(x-1) \right]. $$
(31)

The term in square brackets above may be further simplified using Eq. (28):

$$ \ell_x \widetilde{\alpha}(x) - \ell_{x-1} \widetilde{\alpha}(x-1) = \sum_{j=0}^{\infty} e^{-(R-x+j) \delta } \ell_{R+j} (1 - e^{-\delta}) = \ell_x \widetilde{\alpha}(x) (1 - e^{-\delta}). $$
(32)

Finally, substituting Eq. (32) in Eq. (31) and using Y as defined in Eq. (27) gives

$$ Z = \sum_{x=E}^{R} m_x \ell_x \widetilde{\alpha}(x) (1 - e^{-\delta}) = Y(1-e^{-\delta}). $$
(33)

As in Assumption 1, where the liability Y t is linearized around δ, we may linearize Z t in Eq. (29):

$$ Z_t = Z e^{ -D_Z(\delta_t - \delta) } + e^{-\delta} (\delta_t - \delta)Y_t, $$
(34)

where D Z  =  − Z −1Z/∂δ. Using Eq. (3) from Assumption 1 as well as Eq. (33), we obtain

$$ \frac{Z_t}{Y_t} = \frac{Z}{Y} \frac{Y}{Y_t} e^{-D_Z(\delta_t - \delta) } + e^{-\delta} (\delta_t - \delta) = (1-e^{-\delta}) e^{ -(D_Z - D_Y)(\delta_t - \delta) } + e^{-\delta} (\delta_t - \delta). $$
(35)

The difference between the modified durations of Z and Y, which appears in Eq. (35), is easily derived from Eq. (33). We note that ∂Z/∂δ = (∂Y/∂δ)(1 − e −δ) + Ye −δ. Upon dividing by both sides of Eq. (33), we find that Z −1(∂Z/∂δ) = Y −1(∂Y/∂δ) + (e δ − 1)−1. Hence, D Z  − D Y  =  −(e δ − 1)−1. Substituting this into Eq. (35) results in

$$ \frac{Z_t}{Y_t} = (1-e^{-\delta}) \exp\left[ -\frac{\delta_t - \delta}{e^{\delta}-1} \right] + e^{-\delta} (\delta_t - \delta) \approx (1-e^{-\delta}) \left[ 1 - \frac{\delta_t - \delta}{e^{\delta}-1} \right] + e^{-\delta} (\delta_t - \delta) = 1-e^{-\delta}. $$
(36)

This concludes the proof of the approximation in Eq. (5). \(\square\)

Appendix: 2 Proof of proposition 1

It immediately follows from Eq. (7) and Assumption 2 that

$$ {{\mathbb{E}}} \left[ {F_t} | {{{\mathcal{F}}}_{t-1}} \right] = \left( F_{t-1} + C_{t-1} - Z \right) {{\mathbb{E}}}\left[ {\exp\left( r_t^A - r_t^L + \delta \right)} \right] = e^{\delta}\left( F_{t-1} + C_{t-1} - Z \right). $$
(37)

From the definition of the loss L t in Eq. (8), we therefore obtain

$$ L_t = e^{\delta}(F_{t-1} + C_{t-1} - Z) - F_t. $$
(38)

It is convenient to work in terms of the unfunded liability, \(\overline{F}_t = 1-F_t,\) and replace Z using Eq. (5), yielding \(\overline{F}_t - e^{\delta}\overline{F}_{t-1} = L_t - e^{\delta} C_{t-1}.\) Upon substituting the supplementary contribution from Eq. (9), we get \(\overline{F}_t - e^{\delta} \overline{F}_{t-1} = L_t - e^{\delta} {\varvec{\pi}}' {\bf L}_{t-1}.\) This is a linear difference Eq. in \(\overline{F}_t\) which may be solved using standard techniques. It is easily verified that its solution is

$$ \overline{F}_t = {\varvec{\lambda}}' {\bf L}_t, $$
(39)

where \({\varvec{\lambda}} = (\lambda_1, \ldots, \lambda_p)'\) and λ j  = e (j-1)δ + π j  − ∑ j k=1 e (j-kπ k for \(j \in [1, p]\). Equation (39) shows that the unfunded liability is the accumulated value of unpaid losses.

Now, the main recurrence relation may be obtained by substituting F t from the budget constraint in Eq. (7) directly into the r.h.s. of Eq. (38). Simplifying with the help of Eq. (5) and using the unfunded liability \(\overline{F}_t = 1 - F_t\) results in

$$ L_t = \left[ \exp(r_t^A - r_t^L) - 1 \right] \left[ e^{\delta}(\overline{F}_{t-1}-C_{t-1})- 1 \right]. $$
(40)

Replacing C t from Eq. (9) and \(\overline{F}_t\) from Eq. (39) then yields the recurrence relation in Eq. (10) in Proposition 1. We note that \({\varvec{\beta}} = e^{\delta}({\varvec{\lambda}} - {\varvec{\pi}})\) and that β p  = 0. \(\square\)

Appendix: 3 Proof of proposition 2

Part (i). \({{\mathbb{E}}{L_t} = {\mathbb{E}}\left[ {\varepsilon_t} \right]{\mathbb{E}}\left[ {v_t} \right] = 0}\), by independence of \(\varepsilon_t\) and v t . \({{\mathbb{E}}{L_t}}\) exists \({\Leftrightarrow\, {\mathbb{E}}{|L_t|}<\infty}\) (Feller 1971, p. 136). {L t } is evidently adapted to \({\{\mathcal{F}_t\}}.\) For \(\tau < t,\, \varepsilon_t\) is independent of \({{\mathcal{F}}_\tau}\) and \({{\mathbb{E}} \left[ {L_t} \, | \, {{\mathcal{F}}_\tau} \right] = {\mathbb{E}}\left[ {\varepsilon_t} \right] {\mathbb{E}} \left[ {v_t} \; | \; {\mathcal{F}_\tau} \right] = 0}.\) Hence, {L t } is a martingale difference sequence (m.d.s.) (Davidson, 1994, p. 230). The covariance result is a property of m.d.s.: see e.g. Davidson (1994, Corollary 15.4). It is easily verified: \({{\mathbb{E}}\left[ {L_t L_\tau} \right] = {\mathbb{E}}\left[ {\varepsilon_t v_t L_\tau} \right] = {\mathbb{E}}\left[ {\varepsilon_t} \right] {\mathbb{E}}\left[ {v_t L_\tau} \right] = 0}\) when t > τ.

Part (ii). Given the results in part (i), we need only consider VarL t to establish covariance-stationarity of {L t }. Since \({{\mathbb{E}}{\varepsilon_t} = 0}\) and \({{\mathbb{E}}\left[ {L_t L_\tau} \right]=0}\) for τ ≠ t, it follows that \({{\mathbb{E}}{v_t^2} = \sum \beta_j^2 {\mathbb{E}}{L_{t-j}^2} + 1}.\) Now, \({{\rm Var}{L_t} = {\mathbb{E}}{L_t^2} = {\mathbb{E}}\left[ {\varepsilon_t^2} \right]{\mathbb{E}}\left[ {v_t^2} \right]},\) giving a pth-order linear difference equation: VarL t  = ∑ p j=1 σ2 β 2 j VarL t-j  + σ2. It is well-known that stability depends on the zeros of the characteristic polynomial \(z^{p-1}-\sigma^2 \beta_1^2 z^{p-2} - \cdots - \sigma^2 \beta_p^2\) satisfying |z| < 1. Since the coefficients \(\sigma^2 \beta_j^2,\, j=1, \ldots, p\) are non-negative, a necessary and sufficient condition for stability is that \(\sum \sigma^2 \beta_j^2 = \sigma^2 {\varvec{\beta}}' {\varvec{\beta}} < 1.\) The initial values \(L_l, \ldots, L_{l-p+1}\) are known at some time l < 0 and, letting \(l\rightarrow -\infty,\) all second moment terms containing the initial values \(L_l, \ldots, L_{l-p+1}\) vanish geometrically iff σ2 bb < 1. Finally, from the pth-order linear difference equation in VarL t above, VarL t converges to \(\sigma^2 /\left( 1 - \sigma^2 \sum \beta_j^2 \right).\) \(\square\)

Appendix: 4 proof of proposition 3

First, note that \(L_t=\varepsilon_t v_t\) (Eq. (11a)) with v t independent of \(\varepsilon_t\) and \(\{\varepsilon_t\}\) being i.i.d. (Assumption 2) and hence trivially stationary and ergodic, so that ergodicity and stationarity carry across from {v t } to {L t } and vice-versa.

Part (i). To apply Theorem 1, we first obtain the conditions under which {L t } is irreducible.

{L t } in Eq. (10) is not a bilinear process, according to Definition 1 because k ≠ 0 in Eq. (18). However, using Eqs. (11a) and (b), we can write

$$ v_t = \sum_{j=1}^{p} b_j v_{t-j} \epsilon_{t-j} - 1, $$
(41)
$$ v_t + 1 = \sum_{j=1}^{p} b_j (v_{t-j} + 1) \epsilon_{t-j} - \sum_{j=1}^{p} b_j \epsilon_{t-j}. $$
(42)

This conforms to the class of bilinear processes in Eq. (20) and Theorem 4 is therefore applicable.

The state-space representation of {v t } consists of the state Eq. (13a), with A t and B t defined in Eqs. (14) and (15) respectively, together with the observation equation \(v_t = {\varvec{\beta}}' {\bf L}_t -1.\) In accordance with Theorem 4, the irreducibility of the Markov chain {L t } in Eq. (13a) may be guaranteed by insisting on the condition on the density of \(\varepsilon_t,\) which does not violate Assumption 2, specifically \({{\mathbb{E}}{\varepsilon_t}=0}.\) Controllability of {v t } follows by comparing Eqs. (41) and (20) and noting that there is no “autoregressive” element and no possibility of coincident roots in the polynomials of Eq. (21).

To conclude the application of Theorem 1, we proceed to show that the technical conditions \({{\mathbb{E}}{\log^+\|{\bf A}_t\|} < \infty}\) and \({{\mathbb{E}}{\log^+\|{\bf B}_t\|} < \infty}\) do hold. Using the maximum row sum matrix norm induced by the max vector norm (Horn and Johnson 1985, p. 295), \(\|{\bf A}_t\| = \max(1, |\varepsilon_t|\sum|\beta_j|)\) and \(\|{\bf B}_t\| = |-\varepsilon_t| = |\varepsilon_t|.\) Hence, \(\log^+\|{\bf A}_t\| = \log\left[ \max( 1, \max( 1, |\varepsilon_t|\sum|\beta_j| ) ) \right] = \log^+\left( |\varepsilon_t|\sum|\beta_j| \right).\)

Now, \(0 \leq \log^+(k|\varepsilon_t|) \leq k|\varepsilon_t|,\) for any k ≥ 0, and hence \({{\mathbb{E}}{ \log^+(k|\varepsilon_t|) } \leq {\mathbb{E}}\left[ { k|\varepsilon_t| } \right]}.\) By assumption, \({{\mathbb{E}}{\varepsilon_t} = 0}.\) Therefore, \({{\mathbb{E}}\left[ {k \varepsilon_t} \right]}\) exists and is finite, which is equivalent to \({{\mathbb{E}}{|k \varepsilon_t|} < \infty}\) (Feller 1971, p. 136). Hence, \({{\mathbb{E}}{ \log^+(k|\varepsilon_t|) } < \infty}.\) Replacing k by ∑|β j | yields \({{\mathbb{E}}{\log^+\|{\bf A}_t\|} < \infty}.\) Replacing k by 1 yields \({{\mathbb{E}}{\log^+\|{\bf B}_t\|} < \infty}.\) This may also be confirmed by a direct application of Theorem 5 of Kristensen (2009).

This concludes the proof of part (i).

Part (ii). This is an application of Theorem 2. In Eq. (41), {v t } conforms to the bilinear process in Definition 1 as required in Theorem 2. Theorem 2 is therefore applicable to {v t } with \(\widetilde{{\bf A}}_t = {\bf A}_t\) and \(\widetilde{{\bf B}}_t = {\bf B}_t.\) A straightforward inspection of the proof of Theorem 2 (Pham 1985, Theorem 4.1) shows that, because B t in Eq. (15) does not contain a quadratic in \(\varepsilon_t\) (as would be the case for a general bilinear process under Theorem 2), we need not insist that \({{\mathbb{E}}{\varepsilon_t^4}<\infty}\) and merely require that \({{\mathbb{E}}{\varepsilon_t^2}<\infty}.\)

We claim that Q = Q I p , where Q is the stationary variance of {L t }, given on the r.h.s. of Eq. (12), is a positive solution of the matrix Eq. (19), thereby satisfying the condition of Theorem 2. This may be seen by noting that \({{\mathbb{E}}{{\bf L}_t}={\bf 0}_p}\) from part (i) of Proposition 2 and using the state Eq. (13a) to obtain \({{\rm Var}{{\bf L}_{t+1}} = {\mathbb{E}}\left[ {{\rm Var} \left[ {{\bf L}_{t+1}} \, | \, {\varepsilon_{t+1}} \right]} \right] + {\rm Var}\left[ { {\mathbb{E}} \left[ { {\bf L}_{t+1} } \, | \, { \varepsilon_{t+1} } \right] } \right] = {\mathbb{E}}\left[ { {\bf A}_{t+1} ({\rm Var}{{\bf L}_{t}}) {\bf A}_{t+1}' } \right] + {\rm Var}{{\bf B}_{t+1}}}.\) In the covariance-stationary state, this equation becomes \({{\rm Var}{{\bf L}_{t}} = {\mathbb{E}}\left[ { {\bf A}_{t} ({\rm Var}{{\bf L}_{t}}) {\bf A}_{t}' } \right] + {\rm Var}{{\bf B}_{t}}}\) which corresponds to Eq. (19), and is satisfied by VarL t  = Q I p in accordance with Proposition 2.

This concludes the proof of part (ii).

Part (iii). This is an application of Theorem 3. In Eq. (41), {v t } conforms to the bilinear process in Definition 1 as required in Theorem 3, which is therefore applicable to {v t } with \(\widetilde{{\bf A}}_t = {\bf A}_t\) and \(\widetilde{{\bf B}}_t = {\bf B}_t.\) Note, from Eq. (11a), that \({{\mathbb{E}}\left[ {L_t^{2m}} \right] = {\mathbb{E}}\left[ {\varepsilon_t^{2m}} \right] {\mathbb{E}}\left[ {v^{2m}} \right]}.\)

Part (iv). From Proposition 2, {L t } is a martingale difference sequence (m.d.s.). Therefore, {C t } and \(\{\overline{F}_t\}\) are weighted sums of m.d.s. from Eqs. (9) and (39). Moments of \(C_t,\, \overline{F}_t\) and L t follow from Eqs. (9), (39), (10) respectively. {L t } is serially uncorrelated but {C t } and \(\{\overline{F}_t\}\) have zero autocorrelation at lags p and greater. Parts (iv) and (v) are consistent for lag τ = 0 since, e.g., \(\sum\nolimits_{j=1}^p ({\varvec{\pi}} {\varvec{\pi}}')_{(j,j)} = \hbox{tr}({\varvec{\pi}} {\varvec{\pi}}') = {\varvec{\pi}}' {\varvec{\pi}}.\) \(\square\)

Appendix: 5 Averaging and amortization

Actuarial standards of practice discuss asset value averaging in detail: see for example ASB (2009). A taxonomy of practical methods is presented by Committee on Retirement Systems Research (2001) and described mathematically by Owadally and Haberman (2004b). This results in averaging losses, say over \({n\in{\mathbb{N}}}\) years, with allowance for interest:

$$ L_t^A = \frac{1}{n} \sum_{j=0}^{n-1} e^{j\delta} L_{t-j}. $$
(43)

Smoothed losses are then paid off, usually in equal tranches, over a number of years (say \({m\in{\mathbb{N}}}\)) through the device of amortization (ASB 2007a; McGill et al. 2004). A supplementary contribution is paid at time t by amortizing every previous “averaged” loss that has not yet been paid in full over, say, m years:

$$ C_t = \frac{1}{\alpha_m} \sum_{j=0}^{m-1} L_{t-j}^A, $$
(44)

where \(\alpha_m = 1 + e^{-\delta} + \cdots + e^{-(m-1)\delta}\) is the present value of an annuity over m years.

Working through the double summation in Eqs. (43) and (44), one may express the supplementary contribution in terms of the filter specified in Eq. (9), with p = m + n − 1. Letting \(\overline{p}=\max(m,n)\) and \(\underline {p}=\min(m,n)\) and using some algebra returns the following filter weights π j .

$$ \pi_j = \left\{ \begin{array}{ll} (n\alpha_m)^{-1} e^{(j-1)\delta} \alpha_j & \hbox{if}\, j\in[1,\underline{p}] \\ (n\alpha_m)^{-1} e^{[\min(j,n)-1]\delta} \alpha_{\underline{p}} & \hbox{if}\, j\in[\underline{p},\overline{p}] \\ (n\alpha_m)^{-1} e^{(n-1)\delta} \alpha_{p-j+1} & \hbox{if}\, j\in[\overline{p},p] \\ 0 & \hbox{otherwise}\\ \end{array}\right. $$
(45)

The weights {π j } above do satisfy the requirement that ∑ p j=1 e −(j-1)δ π j  = 1 as set out under Eq. (9). Equation (9) shows that the supplementary contribution includes a proportion π j of the loss that occurred j years ago, \(j\in[1,p].\)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Owadally, I. Tail risk in pension funds: an analysis using ARCH models and bilinear processes. Rev Quant Finan Acc 43, 301–331 (2014). https://doi.org/10.1007/s11156-013-0373-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11156-013-0373-9

Keywords

JEL Classification

Navigation