Skip to main content

Model-Based Measurement of Sector Concentration Risk in Credit Portfolios

  • Chapter
  • First Online:
Risk Management in Credit Portfolios

Part of the book series: Contributions to Economics ((CE))

Abstract

The focus of this chapter is on sector concentrations. This type of concentration risk can occur if there is more than one systematic risk factor that influences credit defaults. The main research questions that are analyzed in this chapter are:

  • How can existing approaches for measuring sector concentration risk be modified and adjusted to be consistent with the Basel framework? Is the risk measure Value at Risk problematic when dealing with sector concentration risk?

  • Which methods are capable of measuring concentration risk and how good do they perform in comparison? What are the advantages and disadvantages of these methods?

In order to deal with these questions, it is initially determined how a multi-factor model can be parameterized to obtain a capital requirement which is consistent with Basel II. Furthermore, the models of Pykhtin (Risk 17(3):85–90, 2004), Cespedes et al. (J Credit Risk, 2(3):57–85, 2006), and Düllmann (Measuring business sector concentration by an infection model. Discussion Paper, Series 2: Banking and Financial Studies, Deutsche Bundesbank, (3), 2006), which have been developed to approximate the risk in the presence of sector concentrations, are presented and modified. Then, the accuracy of these models concerning their ability to measure sector concentration risk is compared.

The main results of this section comply with Gürtler et al. (2010).

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.00
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    BCBS (2005a), § 773.

  2. 2.

    Cf. Sect. 3.3.

  3. 3.

    The multi-factor model of Cespedes et al. (2006) is also specified against the background of the regulatory capital formula. However, within the deriviation of their formulas, the authors assume the regulatory capital requirement to be the upper barrier of risk, which is not consistent with the view of supervisors that we presented in Sect. 3.3 and especially in Fig. 3.2. Cf. Sect. 5.2.3 for details regarding this issue.

  4. 4.

    Cf. Sect. 2.2.3.

  5. 5.

    As shown by Morinaga and Shiina (2005), an assignment of borrowers to the wrong sectors usually leads to a higher estimation error than a non-optimal sector definition.

  6. 6.

    In order to allow for negative intra-sector correlations, the factor loading could also be written as r i instead of \( \sqrt {{{\rho_{{\text{Intra}},i}}}} \). However, it is economically reasonable to assume that there is a positive relationship between the asset return of an obligor and the corresponding industry-sector. Thus, the chosen notation should be no practical limitation.

  7. 7.

    Concretely, the independence of the risk factors is essential for the derivation of the Pykhtin-model in Sect. 5.2.2.

  8. 8.

    This approach is a common mathematical method to generate correlated random variables and leads to the identical number of independent risk factors \( {\tilde{z}_k} \) and dependent sector factors \( {\tilde{x}_s} \), that is K equals S. Another common method to determine independent risk factors is the principal component analysis, which leads to a reduced number of risk factors.

  9. 9.

    An overview of the literature regarding the measurement of asset correlation parameters can be found in Düllmann et al. (2008) and Grundke (2008).

  10. 10.

    The correlation structure based on the MSCI US is similar, see Düllmann and Masschelein (2007).

  11. 11.

    Düllmann and Masschelein (2007) notice that the concentration is very similar to other countries like France, Belgium, and Spain.

  12. 12.

    This value results on the basis of both measures (VaR and ES) at the respective confidence level as described in Sect. 4.3.1. The result is consistent with Düllmann and Masschelein (2007), who use a constant intra-sector correlation of 25% in their analysis.

  13. 13.

    See Fig. 4.7 for the portfolio characteristics.

  14. 14.

    We tried several different functional forms but the formula above performed best. The multipliers 18.5% and 34% in function (5.8) were determined with a grid search using a reasonable parameter range, which is similar to the procedure of Lopez (2004) used for the single correlation parameter.

  15. 15.

    E.g. Heitfield et al. (2006) determine the sector loadings, which equal \( \sqrt {{{\rho_{\text{Intra}}}}} \), for 50 industry sectors using KMV data on asset values. The resulting intra-sector correlation is on average 18.8% and the standard deviation is 8.3%. These inter-sectoral differences are not captured by the formula above.

  16. 16.

    A correlation structure with one degree of freedom for every PD/sector-combination is practically unfeasible due to high data requirements.

  17. 17.

    In our setting, the computation time could be reduced by more than 99.8%.

  18. 18.

    The conditional PD stems from the Vasicek model, cf. Sect. 2.4 or 2.7.

  19. 19.

    Cf. (5.9).

  20. 20.

    For the determination of c i , we need both the intra- and inter-sector correlations, which can be taken from Sect. 5.2.1.

  21. 21.

    This formula has already been derived for the granularity adjustment formula, cf. (4.18).

  22. 22.

    Cf. Sect. 4.2.1.2.

  23. 23.

    Cf. (5.2).

  24. 24.

    See Appendix 5.5.2.

  25. 25.

    The derivation of the variance decomposition can be found in Weiss (2005), p. 385 f.

  26. 26.

    The quadratic computation effort is due to the determination of a double sum (see (5.22) and (5.24)).

  27. 27.

    The results of the multi-factor adjustment do not differ whether different exposures with the same PD are aggregated or handled separately on borrower level. For details see Sect. 5.2.2.1 and Appendix 5.5.1.

  28. 28.

    The computation time when calculating the multi-factor adjustment on bucket- instead on borrower-level can be reduced from 67 min to 5 s for a portfolio with 11 sectors, 7 PD-classes, and 5,000 creditors.

  29. 29.

    In the strict sense, Cespedes et al. (2006) relate the multi-factor model to the economic capital in a single-factor model. But since they apply the regulatory capital formula and we require a relation to this formula, too, we use the term regulatory capital instead.

  30. 30.

    The idea is related to Pykhtin (2004), who uses the VaR from the ASRF model as a weight when maximizing the correlation between the single factor of the comparable one-factor model and the sector factors; cf. (5.82)–(5.85).

  31. 31.

    Cespedes et al. (2006) call this parameter the capital diversification index (CDI).

  32. 32.

    This concentration measure corresponds to (2.87).

  33. 33.

    The setting is similar to Cespedes et al. (2006). Until this point, the main difference is the definition of the intra- and inter-sector correlations.

  34. 34.

    For the determination of the economic capital for one specific portfolio, the number of trials is slightly low but as we perform 25,000 simulations and the simulation noise of each simulation is unsystematic, the error terms should cancel out each other to a large extent.

  35. 35.

    We have also tested the results when using the ES instead of the unexpected loss but the coefficient of determination is higher when subtracting the EL in the corresponding formulas when performing the simulations.

  36. 36.

    To determine the Expected Shortfall with (4.59), a bivariate cumulative normal distribution has to be computed whereas the Value at Risk only makes use of univariate distributions.

  37. 37.

    We have tried several different regressions but similar to Cespedes et al. (2006), this function worked best. In contrast to Cespedes et al. (2006) we do not set the first parameter a 0 to one because our DF-factor is not bounded by the single-factor model.

  38. 38.

    The shape of the function is similar to Cespedes et al. (2006) but their range is from 0.1 to 1.0 whereas our function ranges from 0.2 to 1.5. In addition, they received a little higher R2 (99.4% instead of 95.5% or 97.9%) but this is mainly due to the different simulation setting. Cespedes et al. (2006) directly draw the parameter \( \bar{\beta } \) as an input parameter for each simulation, implying \( \bar{\beta } \) to fully define their correlation structure. We use a heterogeneous correlation structure instead and compute \( \bar{\beta } \) for the portfolios. Thus, in our setting \( \bar{\beta } \) does not reflect the complete correlation structure, which results in a lower R2 but does not imply a worse approximation.

  39. 39.

    In comparison to the original procedure, the computation time could be reduced by almost 99.9% in our calculations.

  40. 40.

    Cf. Cifuentes et al. (1996), Cifuentes and O’Connor (1996), and Cifuentes and Wilcox (1998).

  41. 41.

    See Appendix 5.5.5.

  42. 42.

    See Appendix 5.5.5.

  43. 43.

    See Appendix 5.5.6.

  44. 44.

    Similar to the BET-model, the authors developed their model for CDOs but it can also be applied to standard credit portfolios.

  45. 45.

    The expected number of defaults in the infectious defaults model is determined in Appendix 5.5.7.

  46. 46.

    Cf. (5.61) for the corresponding expression without using the parameters of the BET-model.

  47. 47.

    Düllmann (2006) mentions that the simultaneous computation of both parameters leads to numerical problems. For this reason, the discrepancy in the EL is accepted. Cf. Düllmann (2006), p. 10.

  48. 48.

    See also definition (5.4) of Sect. 5.2.1.

  49. 49.

    The portfolios used for calibration correspond to the setting of Düllmann (2006).

  50. 50.

    Due to the characteristic of the correlation parameter (5.67), the parameter \( {\bar{r}_{\text{Inter}}} \) is always smaller than the parameter \( {\bar{r}_{\text{Intra}}} \).

  51. 51.

    See (5.47).

  52. 52.

    For the second fraction, a number of n elements has to be computed. Depending on the number of buckets or credits, the computation time can be longer than for the first term, but due to the linearity this term is virtually unproblematic.

  53. 53.

    The computation time when calculating the infection model on bucket- instead on borrower-level can be reduced from 12 min to less than 1 s for a portfolio with 11 sectors, 7 PD-classes, and 5,000 creditors.

  54. 54.

    The results refer to the total gross loss of a portfolio in terms of ES or VaR. To relate this to the unexpected net loss, the results have to be multiplied by the LGD and the EL has to be subtracted.

  55. 55.

    The small mismatch is mainly due to keeping the ES-confidence level constant and not a result of the chosen intra-sector correlation function. If we directly compare the results from Monte Carlo simulations with the ES in the ASRF framework, the relative root mean squared error is reduced from 0.97% to 0.28%.

  56. 56.

    In our analyses, the number of simulation runs is 500,000.

  57. 57.

    If we consider all 25,000 simulated portfolios from Sect. 5.2.3, the lowest measured economic capital requirement was even 26% lower than the regulatory capital. This underlines the prospects of actively managing credit portfolios, e.g. with credit derivatives, but this is not in the scope of this thesis.

  58. 58.

    CHKR I still corresponds to the DF-function based on Monte Carlo simulation and CHKR II on the Pykhtin formula.

  59. 59.

    Acharya et al. (2006) examined credit portfolios of 105 Italian banks during the period 1993–1999. In this study, most bank portfolios had a HHI between 20% and 30%. However, it has to be considered that the number of different industry sectors was 23 whereas we use 11 different sectors. Thus, for a comparable degree of diversification their calculated HHI have to be slightly smaller than our HHIs.

  60. 60.

    The runtimes refer to a quad-core PC with 2.66 GHz CPUs (calculated on one core).

  61. 61.

    In contrast to this representation, Pykhtin (2004) applies these and the following formulas to n sector factors whereas we use K sector factors with \( \tilde{\bar{x}} \). This can lead to a significant reduction of the computation time as will be shown later on.

  62. 62.

    This simplification of the Taylor series could already be used for the granularity adjustment in Sect. 4.2.1.1.

  63. 63.

    Cf. Pykhtin (2004).

  64. 64.

    The following calculations are based on Tasche (2006a), p. 41 ff.

  65. 65.

    It has to be noticed that the conditional correlation matrix is symmetric, so we have \( \begin{array} {c} \frac{{d\eta_{2,c}^\infty \left( {\bar{x}} \right)}}{{d\bar{x}}} = \sum\limits_{i = 1}^n {\sum\limits_{j = 1}^n {{w_i}} } {w_j}ELG{D_i}ELG{D_j} \\ \cdot \left( {\frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_j}\left( {\bar{x}} \right)} \right] - \rho_{ij}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)} \right. \\ + \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right] - \rho_{ij}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_j}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) \\ \left. { - \frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}{p_j}\left( {\bar{x}} \right) - \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}}{p_i}\left( {\bar{x}} \right)} \right). \\ \end{array} \) for all i, j.

References

  • Acharya VV, Hasan I, Saunders A (2006) Should banks be diversified? Evidence from individual bank loan portfolios. J Bus 79(3):1355–1412

    Article  Google Scholar 

  • Basel Committee on Banking Supervision (2005a) International convergence of capital measurement and capital standards – a revised framework, Updated November 2005, Bank for International Settlements, Basel

    Google Scholar 

  • Basel Committee on Banking Supervision (2006) Studies on credit risk concentration: an overview of the issues and a synopsis of the results from the research task force project. BCBS Working Paper No. 15, Bank for International Settlements, Basel

    Google Scholar 

  • Cespedes J, de Juan Herrero J, Kreinin A, Rosen D (2006) A simple multi-factor “factor adjustment” for the treatment of diversification in credit capital rules. J Credit Risk 2(3):57–85

    Google Scholar 

  • Cifuentes A, O’Connor G (1996) The binomial expansion method applied to CBO/CLO analysis, Special report. Moody’s Investor Service, New York

    Google Scholar 

  • Cifuentes A, Wilcox C (1998) The double binomial method and its application to a special case of CDO structures, Special report. Moody’s Investor Service, New York

    Google Scholar 

  • Cifuentes A, Murphy E, O’Connor G (1996) Emerging market collateralized bond obligations: an overview, Special report. Moody’s Investor Service, New York

    Google Scholar 

  • Davis M, Lo V (2001) Infectious defaults. Quant Fin 1(4):382–387

    Article  Google Scholar 

  • Düllmann K (2006). Measuring business sector concentration by an infection model. Discussion Paper, Series 2: Banking and financial studies. Deutsche Bundesbank (3)

    Google Scholar 

  • Düllmann K (2007) Measuring concentration risk in credit portfolios. In: Christodoulakis G, Satchell S (eds) The analytics of risk model validation. Academic Press, Amsterdam, pp 59–78

    Google Scholar 

  • Düllmann K, Masschelein N (2007) A tractable model to measure sector concentration risk in credit portfolios. J Fin Serv Res 32(1):55–79

    Article  Google Scholar 

  • Düllmann K, Küll J, Kunisch M (2008) Estimating asset correlations from stock prices or default rates – Which method is superior? Discussion Paper, Series 2: Banking and financial studies, Deutsche Bundesbank (4)

    Google Scholar 

  • Gordy MB (2003) A risk-factor model foundation for rating-based capital rules. J Fin Intermediation 12(3):199–232

    Article  Google Scholar 

  • Grundke P (2008) Regulatory treatment of the double default effect under the New Basel accord: how conservative is it? Rev Manag Sci 2(1):37–59

    Article  Google Scholar 

  • Gürtler M, Hibbeln M, Vöhringer C (2010) Measuring concentration risk for regulatory purposes. J Risk 12(3):69–104

    Google Scholar 

  • Heitfield E, Burton S, Chomsisengphet S (2006) Systematic and idiosyncratic risk in syndicated loan portfolios. J Credit Risk 2(3):3–31

    Google Scholar 

  • Lopez J (2004) The empirical relationship between average asset correlation, firm probability of default, and asset size. J Fin Intermediation 13(2):265–283

    Article  Google Scholar 

  • Martin R, Wilde T (2002) Unsystematic credit risk. Risk 15(11):123–128

    Google Scholar 

  • Morinaga S, Shiina Y (2005) Underestimation of sector concentration risk by mis-assignment of borrowers. Working Paper, Tokyo

    Google Scholar 

  • Pykhtin M (2004) Multi-factor adjustment. Risk 17(3):85–90

    Google Scholar 

  • Tasche D (2006a) Anwendungen der Stochastik in der Bankenaufsicht. Lecture Notes, Frankfurt University. http://www-m4.ma.tum.de/pers/tasche. Accessed 18 Aug 2009

  • Weiss NA (2005) A course in probability. Addison-Wesley, Boston

    Google Scholar 

  • Wilde T (2001) Probing granularity. Risk 14(8):103–106

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Martin Hibbeln .

Appendix

Appendix

5.1.1 Optimal Choice of the Single Correlation Factor

To relate \( 1/11 = 9.1\% \) to \( \tilde{\bar{L}} \), it is assumed that the new systematic factor \( \tilde{L} \) has a linear dependence to the original sector factors:Footnote 61

$$ K \leq n $$
(5.73)
$$ \tilde{\bar{x}} = \sum\limits_{k = 1}^K {{b_k} \cdot {{\tilde{z}}_k},} $$
(5.74)

Condition (5.74) satisfies that the new systematic factor still has a variance of 1. In order to specify the correlation factors c i and the coefficients b k , it will be required that the loss \( {\text{with}}\,\,\,\sum\limits_{k = 1}^K {b_k^2 = 1.} \) equals the conditional expectation of the “true” loss \( \tilde{\bar{L}} \). This assures that the first element of the subsequently performed Taylor series expansion vanishes.Footnote 62 To determine \( b{E}(\tilde{L}|\tilde{\bar{x}}) \), we first recall that the asset return of obligor i in sector s can be written as

$$ b{E}(\tilde{\bar{L}}|\tilde{\bar{x}}) $$
(5.75)

Now, each original sector factor \( {\tilde{a}_{s,i}} = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\tilde{x}_s} + \sqrt {{1 - {\rho_{{\text{Intra}},i}}}} \cdot {\tilde{\xi }_i}. \) is decomposed into a part that is related to the single-factor \( {\tilde{x}_s} \) and a part that is independent of this factor:

$$ \tilde{\bar{x}} $$
(5.76)

with \( {\tilde{x}_s} = {\bar{\rho }_s} \cdot \tilde{\bar{x}} + \sqrt {{1 - {{\bar{\rho }}_s}^2}} \cdot {\tilde{\eta }_s} \). Using (5.2), (5.73), and the independence of \( {\tilde{\eta }_s} \sim \mathcal{N}(0,1) \) if \( {\tilde{z}_i},\;{\tilde{z}_j} \), the correlation parameter \( i \ne j \) can be expressed as

$$ {\bar{\rho }_s} $$
(5.77)

Using this notation, the asset return (5.75) can now be written as

$$ \begin{array} {c} {{\bar{\rho }}_s} = {\text{Corr}}\left( {{{\tilde{x}}_s},\tilde{\bar{x}}} \right) = {\text{Corr}}\left( {\sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {{\tilde{z}}_k}}, \sum\limits_{k = 1}^K {{b_k} \cdot {{\tilde{z}}_k}} } \right) \\ = \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {b_k} \cdot b{V}\left( {{{\tilde{z}}_k}} \right)} = \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {b_k}.} \\ \end{array} $$
(5.78)

The independent standard normally distributed random variables \( \begin{array} {c} {{\tilde{a}}_{s,i}} = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\tilde{x}}_s} + \sqrt {{1 - {\rho_{{\text{Intra}},i}}}} \cdot {{\tilde{\xi }}_i} \\ = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \left( {{{\bar{\rho }}_s} \cdot \tilde{\bar{x}} + \sqrt {{1 - {{\bar{\rho }}_s}^2}} \cdot {{\tilde{\eta }}_s}} \right) + \sqrt {{1 - {\rho_{{\text{Intra}},i}}}} \cdot {{\tilde{\xi }}_i} \\ = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_s} \cdot \tilde{\bar{x}} + \sqrt {{{\rho_{{\text{Intra}},i}} - {\rho_{{\text{Intra}},i}} \cdot {{\bar{\rho }}_s}^2}} \cdot {{\tilde{\eta }}_s} + \sqrt {{1 - {\rho_{{\text{Intra}},i}}}} \cdot {{\tilde{\xi }}_i}. \\ \end{array} \) and \( {\tilde{\eta }_s} \) can be combined into a new standard normally distributed random variable \( {\tilde{\xi }_i} \), leading to

$$ {\tilde{\zeta }_i} $$
(5.79)

with \( {\tilde{a}_{s,i}} = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\bar{\rho }_i} \cdot \tilde{\bar{x}} + \sqrt {{1 - {{\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i}} \right)}^2}}} \cdot {\tilde{\zeta }_i}, \) for each obligor i in sector s. Since the variable \( {\bar{\rho }_i} = {\bar{\rho }_s} \) is independent of \( {\tilde{\zeta }_i} \), we can use the known formula of the single-factor model for the conditional expectation

$$ \tilde{\bar{x}} $$
(5.80)

The mentioned condition \( b{E}\left( {\tilde{L}|\tilde{\bar{x}}} \right) = \sum\limits_{i = 1}^n {{w_i} \cdot LG{D_i} \cdot \Phi \left[ {\frac{{{\Phi^{ - 1}}(P{D_i}) - \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i} \cdot {\Phi^{ - 1}}(\alpha )}}{{\sqrt {{1 - {{\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i}} \right)}^2}}} }}} \right]} . \) leads to

$$ \tilde{\bar{L}} = b{E}\left( {\tilde{L}|\tilde{\bar{x}}} \right) $$
(5.81)

using (5.9), (5.80), (5.77), and \( \begin{array} {lll} \tilde{\bar{L}} &= b{E}\left( {\tilde{L}|\tilde{\bar{x}}} \right) \\ & \Leftrightarrow \Phi \left[ {\frac{{{\Phi^{ - 1}}(P{D_i}) - {c_i} \cdot \tilde{\bar{x}}}}{{\sqrt {{1 - {c_i}^2}} }}} \right] = \Phi \left[ {\frac{{{\Phi^{ - 1}}(P{D_i}) - \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i} \cdot {\Phi^{ - 1}}(\alpha )}}{{\sqrt {{1 - {{\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i}} \right)}^2}}} }}} \right] \\ & \Leftrightarrow {c_i} = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {{\bar{\rho }}_i} = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \sum\limits_{k = 1}^K {{\alpha_{i,k}} \cdot {b_k}}, \\ \end{array} \) for each obligor i in sector s. While \( {\alpha_{i,k}} = {\alpha_{s,k}} \) and \( {\rho_{{\text{Intra}},i}} \) are known, the coefficients \( {\alpha_{i,k}} \) are unknown.

While (5.81) already satisfies that the first-order term of the Taylor series vanishes, the concrete choice of the parameter set {\( {b_k} \)} is critical concerning the distance between the zeroth-order term \( {b_k} \) and the unknown quantile \( {q_\alpha }(\tilde{\bar{L}}) \). Unfortunately, it is not obvious how this distance can be minimized. Thus, Pykhtin (2004) relies on the intuition that coefficients which maximize the (weighted) correlation between the single factor \( {q_\alpha }(\tilde{L}) \) and the sector factors {\( \tilde{\bar{x}} \)} should lead to good results. This leads to the following maximization problem:

$$ {\tilde{x}_s} $$
(5.82)

subject to

$$ \mathop {{\max }}\limits_{\left\{ {{b_k}} \right\}} \left( {\sum\limits_{i = 1}^n {{d_i} \cdot {{\bar{\rho }}_i}} } \right) = \mathop {{\max }}\limits_{\left\{ {{b_k}} \right\}} \left( {\sum\limits_{i = 1}^n {{d_i} \cdot \sum\limits_{k = 1}^K {{\alpha_{i,k}} \cdot {b_k}} } } \right) $$
(5.83)

The solution of this optimization problem isFootnote 63

$$ \sum\limits_{k = 1}^K {b_k^2 = 1.} $$
(5.84)

where the positive constant Lagrange multiplier \( {b_k} = \sum\limits_{i = 1}^n {\frac{{{d_i} \cdot {\alpha_{ik}}}}{{2\tau }}}, \) is chosen in a way that {\( \tau \)} satisfies the constraint. As a final step, the weighting factor d i has to be chosen. After trying several specifications, Pykhtin (2004) uses

$$ {b_k} $$
(5.85)

which is the VaR formula in a single-factor model. The intuition behind this choice is that obligors with a high exposure in terms of VaR should have a large weight in the maximization problem whereas obligors with a small VaR should have a minor impact. Summing up, the correlation parameter c i results from (5.81), where the coefficients b k are determined by (5.83)–(5.85).

5.1.2 Conditional Correlation

The correlation conditional on \( {d_i} = {w_i} \cdot LG{D_i} \cdot \Phi \left[ {\frac{{{\Phi^{ - 1}}(P{D_i}) + \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\Phi^{ - 1}}(\alpha )}}{{\sqrt {{1 - {\rho_{{\text{Intra}},i}}}} }}} \right], \) between the asset returns from (5.19) can be written as

$$ \tilde{\bar{x}} $$
(5.86)

using the independence of the factors \( \begin{array} {c} \rho_{ij}^{\bar{x}} = {\text{Corr}}\left( {{{\tilde{a}}_{s,i}},{{\tilde{a}}_{t,j}}|\tilde{\bar{x}}} \right) \\ = \frac{{{\text{Cov}}\left( {\sum\limits_{k = 1}^K {\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\alpha_{s,k}} - {c_i} \cdot {b_k}} \right)} \cdot {{\tilde{z}}_k},\sum\limits_{k = 1}^K {\left( {\sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot {\alpha_{t,k}} - {c_j} \cdot {b_k}} \right)} \cdot {{\tilde{z}}_k}} \right)}}{{\sqrt {{b{V}\left( {{{\tilde{a}}_{s,i}}|\tilde{\bar{x}}} \right)}} \cdot \sqrt {{b{V}\left( {{{\tilde{a}}_{t,j}}|\tilde{\bar{x}}} \right)}} }} \\ = \frac{{\sum\limits_{k = 1}^K {\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\alpha_{s,k}} - {c_i} \cdot {b_k}} \right) \cdot \left( {\sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot {\alpha_{t,k}} - {c_j} \cdot {b_k}} \right)} }}{{\sqrt {{1 - {c_i}^2}} \cdot \sqrt {{1 - {c_j}^2}} }}, \\ \end{array} \). The numerator can be simplified using \( {\tilde{z}_k} \) from (5.81) and \( \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {b_k}} = {{{c_i}} \mathord{\left/{\vphantom {{{c_i}} {\sqrt {{{\rho_{{\text{Intra}},i}}}} }}} \right.} {\sqrt {{{\rho_{{\text{Intra}},i}}}} }} \) from (5.74):

$$ \sum\limits_{k = 1}^K {b_k^2 = 1} $$
(5.87)

This leads to

$$ \begin{array} {lll} \sum\limits_{k = 1}^K {\left( {\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {\alpha_{s,k}} - {c_i} \cdot {b_k}} \right) \cdot \left( {\sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot {\alpha_{t,k}} - {c_j} \cdot {b_k}} \right)} \\ = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {\alpha_{t,k}}} - \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {c_j} \cdot \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {b_k}} \\ & - \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot {c_i} \cdot \sum\limits_{k = 1}^K {{\alpha_{t,k}} \cdot {b_k}} + {c_i} \cdot {c_j} \cdot \sum\limits_{k = 1}^K {{b_k}^2} \\ = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {\alpha_{t,k}}} - \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot {c_j} \cdot \frac{{{c_i}}}{{\sqrt {{{\rho_{{\text{Intra}},i}}}} }} \\ & - \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot {c_i} \cdot \frac{{{c_j}}}{{\sqrt {{{\rho_{{\text{Intra}},j}}}} }} + {c_i} \cdot {c_j} \\ = \sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {\alpha_{t,k}}} - {c_j} \cdot {c_i}. \\ \end{array} $$
(5.88)

5.1.3 Calculation of the Decomposed Variance

In order to determine the conditional variance, it is decomposed into the following terms:Footnote 64

$$ \rho_{ij}^{\bar{x}} = \frac{{\sqrt {{{\rho_{{\text{Intra}},i}}}} \cdot \sqrt {{{\rho_{{\text{Intra}},j}}}} \cdot \sum\limits_{k = 1}^K {{\alpha_{s,k}} \cdot {\alpha_{t,k}}} - {c_i} \cdot {c_j}}}{{\sqrt {{1 - {c_i}^2}} \cdot \sqrt {{1 - {c_j}^2}} }}. $$
(5.89)

For calculation of these terms, first the expressions (a) \( b{V}\left( {\tilde{L}|\tilde{\bar{x}} = \bar{x}} \right) = b{V}\left[ {b{E}\left( {\tilde{L}|\left\{ {{{\tilde{z}}_k}} \right\}} \right)|\tilde{\bar{x}} = \bar{x}} \right] + b{E}\left[ {b{V}\left( {\tilde{L}|\left\{ {{{\tilde{z}}_k}} \right\}} \right)|\tilde{\bar{x}} = \bar{x}} \right] \), (b) \( b{E}(\tilde{L}|\{ {\tilde{z}_k}\} ) \), and (c) \( b{E}({\tilde{L}^2}|\{ {\tilde{z}_k}\} ) \) will be calculated. The conditional loss is given as

$$ b{V}(\tilde{L}|\{ {\tilde{z}_k}\} ) $$
(5.90)

and for stochastically independent LGDs this leads to

$$ \tilde{L}|\{ {\tilde{z}_k}\} = \sum\limits_i {{w_i}} \cdot \left( {{{\widetilde{{LGD}}}_i}|\{ {{\tilde{z}}_k}\} } \right) \cdot \left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right), $$
(5.91)
  1. (a)

    With \( \tilde{L}|\{ {\tilde{z}_k}\} = \sum\limits_i {{w_i}} \cdot {\widetilde{{LGD}}_i} \cdot \left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right). \) and \( b{E}\left( {{{\widetilde{{LGD}}}_i}} \right) = :ELG{D_i} \) we obtain:

$$ b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right) = :{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) $$
(5.92)
  1. (b)

    Consider that \( b{E}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right) = \sum\limits_i {{w_i}} \cdot ELG{D_i} \cdot {p_i}\left( {\{ {{\tilde{z}}_k}\} } \right). \), \( 1_{\left\{ {{{\tilde{D}}_i}} \right\}}^2 = {1_{\left\{ {{{\tilde{D}}_i}} \right\}}} \), and

$$ b{E}\left( {{{\widetilde{{LGD}}}^2}} \right) = {b{E}^2}\left( {\widetilde{{LGD}}} \right) + b{V}\left( {\widetilde{{LGD}}} \right) = :ELG{D^2} + VLGD $$
(5.93)

as well as

$$ \begin{array} {c} b{E}\left( {LG{D_i}LG{D_j}} \right) = {\text{Cov}}\left( {LG{D_i},LG{D_j}} \right) + b{E}\left( {LG{D_i}} \right)b{E}\left( {LG{D_j}} \right) \\ = b{E}\left( {LG{D_i}} \right)b{E}\left( {LG{D_j}} \right) \\ = :ELG{D_i}ELG{D_j}, \\ \end{array} $$
(5.94)

Moreover, we have

$$ \begin{array} {c} b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right) = {\text{Cov}}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}},{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right) + b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right)b{E}\left( {{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right) \\ = b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right)b{E}\left( {{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right) \\ = {p_i}\left( {\{ {{\tilde{z}}_k}\} } \right){p_j}\left( {\{ {{\tilde{z}}_k}\} } \right). \\ \end{array} $$
(5.95)
$$ {\left( {\sum\limits_i {{x_i}} } \right)^2} = \sum\limits_i {\sum\limits_j {{x_i}{x_j} = \sum\limits_i {{x_i}^2} + \sum\limits_i {\sum\limits_{j \ne i} {{x_i}{x_j}} } } } $$
(5.96)

Thus, we obtain:

$$ \sum\limits_{j \ne i} {{x_i}{x_j}} = \sum\limits_j {{x_i}{x_j}} - {x_i}^2. $$
(5.97)
  1. (c)

    The conditional variance \( \begin{array} {c} b{E}\left( {{{\tilde{L}}^2}|\{ {{\tilde{z}}_k}\} } \right) = b{E}\left[ {{{\sum\limits_i {\left( {{w_i}{{\widetilde{{LGD}}}_i}{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} \right)} }^2}|\{ {{\tilde{z}}_k}\} } \right] \\ = b{E}\left[ {\sum\limits_i {{w_i}^2} {{\widetilde{{LGD}}}_i}^21_{\left\{ {{{\tilde{D}}_i}} \right\}}^2|\{ {{\tilde{z}}_k}\} } \right] + b{E}\left[ {\sum\limits_i {\sum\limits_{j \ne i} {{w_i}{w_j}{{\widetilde{{LGD}}}_i}{{\widetilde{{LGD}}}_j}{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}|\{ {{\tilde{z}}_k}\} } } } \right] \\ = \sum\limits_i {{w_i}^2} b{E}\left( {LG{D_i}^2} \right){p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) \\ + \sum\limits_i {\sum\limits_{j \ne i} {{w_i}{w_j}b{E}\left( {LG{D_i}LG{D_j}} \right)b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}{1_{\left\{ {{{\tilde{D}}_j}} \right\}}}} \right)} } \\ = \sum\limits_i {{w_i}^2} \left( {ELG{D_i}^2 + VLG{D_i}} \right){p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) \\ + \sum\limits_i {\sum\limits_{j \ne i} {{w_i}{w_j}ELG{D_i}ELG{D_j}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right){p_j}\left( {\{ {{\tilde{z}}_k}\} } \right)} } \\ = \sum\limits_i {{w_i}^2} \left( {ELG{D_i}^2 + VLG{D_i}} \right){p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) - \sum\limits_i {{w_i}^2ELG{D_i}^2{p_i}^2\left( {\{ {{\tilde{z}}_k}\} } \right)} \\ + \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right){p_j}\left( {\{ {{\tilde{z}}_k}\} } \right)} } \\ = \sum\limits_i {{w_i}^2} \left( {ELG{D_i}^2\left[ {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) - {p_i}^2\left( {\{ {{\tilde{z}}_k}\} } \right)} \right] + VLG{D_i}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)} \right) \\ + {b{E}^2}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right). \\ \end{array} \) is equal to

$$ b{V}(\tilde{L}|\{ {\tilde{z}_k}\} ) $$
(5.98)
  1. (d)

    Using the law of iterated expectation, we have

$$ \begin{array} {c} b{V}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right) = b{E}\left( {{{\tilde{L}}^2}|\{ {{\tilde{z}}_k}\} } \right) - {b{E}^2}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right) \\ = \sum\limits_i {{w_i}^2} \left( {ELG{D_i}^2\left[ {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) - {p_i}^2\left( {\{ {{\tilde{z}}_k}\} } \right)} \right] + VLG{D_i}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)} \right). \\ \end{array} $$
(5.99)

Thus, with (5.98) the expectation of the conditional variance can be written as

$$ {p_i}(\bar{x}) = b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\tilde{\bar{x}} = \bar{x}} \right) = b{E}\left[ {b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}|\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right] = b{E}\left[ {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right] $$
(5.100)

For independent idiosyncratic factors \( \begin{array} {c} b{E}\left[ {b{V}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\tilde{\bar{x}} = \bar{x}} \right] = \sum\limits_i {{w_i}^2\left( {ELG{D_i}^2\left( {b{E}\left[ {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right] - b{E}\left[ {{p_i}^2\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right]} \right)} \right.} \\ \left. { + VLG{D_i}b{E}\left[ {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right]} \right) \\ = \sum\limits_i {{w_i}^2\left( {ELG{D_i}^2\left( {{p_i}(\bar{x}) - b{P}\left[ {\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}} = 1} \right) \wedge \left( {{1_{\left\{ {{{\tilde{D}}_i}^\prime } \right\}}} = 1} \right)|\bar{x}} \right]} \right)} \right.} \\ \left. { + VLG{D_i}{p_i}(\bar{x})} \right). \\ \end{array} \) and with

$$ {\tilde{\zeta }_i},{\tilde{\zeta }_i}^\prime \sim \mathcal{N}(0,1) $$
(5.101)

we get

$$ {p_i}(\bar{x}): = \Phi \left( {\frac{{{\Phi^{ - 1}}(P{D_i}) - {c_i} \cdot \bar{x}}}{{\sqrt {{1 - {c_i}^2}} }}} \right) \Leftrightarrow \frac{{{\Phi^{ - 1}}(PD) - {c_i} \cdot \bar{x}}}{{\sqrt {{1 - {c_i}^2}} }} = {\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right), $$
(5.102)

with the correlation conditional on \( \begin{array} {c} b{P}\left[ {\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}} = 1} \right) \wedge \left( {{1_{\left\{ {{{\tilde{D}}_i}^\prime } \right\}}} = 1} \right)|\bar{x}} \right] \\ = b{P}\left[ {{c_i} \cdot \tilde{\bar{x}} + \sqrt {{1 - {c_i}^2}} \cdot {{\tilde{\zeta }}_i} \leq {\Phi^{ - 1}}(P{D_i}),{c_i} \cdot \tilde{x} + \sqrt {{1 - {c_i}^2}} \cdot {{\tilde{\zeta }}_i}^\prime \leq {\Phi^{ - 1}}(P{D_i})|\bar{x}} \right] \\ = b{P}\left[ {{{\tilde{\zeta }}_i} \leq \frac{{{\Phi^{ - 1}}(P{D_i}) - {c_i} \cdot \bar{x}}}{{\sqrt {{1 - {c_i}^2}} }},{{\tilde{\zeta }}_i}^\prime \leq \frac{{{\Phi^{ - 1}}(P{D_i}) - {c_i} \cdot \bar{x}}}{{\sqrt {{1 - {c_i}^2}} }}} \right] \\ = {\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right),{\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right),\rho_{ii}^{\bar{x}}} \right), \\ \end{array} \) of (5.20). Hence, (5.100) results in

$$ \bar{x} $$
(5.103)
  1. (e)

    Using (5.92), the variance of the conditional expectation can be expressed as

$$ \begin{array} {c} b{E}\left[ {b{V}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\tilde{\bar{x}} = \bar{x}} \right] = \sum\limits_i {{w_i}^2\left( {ELG{D_i}^2\left[ {{p_i}(\bar{x}) - {\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right),{\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right),\rho_{ii}^{\bar{x}}} \right)} \right]} \right.} \\ \left. { + VLG{D_i}{p_i}(\bar{x})} \right). \\ \end{array} $$
(5.104)

leading to

$$ \begin{array} {c} b{V}\left[ {b{E}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\tilde{\bar{x}} = x} \right] \\ = b{E}\left[ {{b{E}^2}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right] - {b{E}^2}\left[ {b{E}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right] \\ = b{E}\left( {{b{E}^2}\left[ {\sum\limits_i {{w_i}{{\widetilde{{LGD}}}_i}{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} |\{ {{\tilde{z}}_k}\} } \right]|\bar{x}} \right) - {b{E}^2}\left( {\sum\limits_i {{w_i}ELG{D_i}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} } \right) \\ = b{E}\left[ {{{\left( {\sum\limits_i {{w_i}ELG{D_i}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right)} } \right)}^2}|\bar{x}} \right] - {\left( {\sum\limits_i {{w_i}ELG{D_i}{p_i}(\bar{x})} } \right)^2}, \\ \end{array} $$
(5.105)

Analogous to (5.102) and using the conditional correlation (5.20), this can be expressed as:

$$ \begin{array} {llll} b{V}\left[ {b{E}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\tilde{\bar{x}} = x} \right] \\ = b{E}\left[ {\sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) \cdot {p_j}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} } } \right] \\ & - \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}{p_i}(\bar{x}){p_j}(\bar{x})} } \\ = \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}b{E}\left( {{p_i}\left( {\{ {{\tilde{z}}_k}\} } \right) \cdot {p_j}\left( {\{ {{\tilde{z}}_k}\} } \right)|\bar{x}} \right)} } \\ & - \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}{p_i}(\bar{x}){p_j}(\bar{x})} } \\ = \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}\left[ {b{P}\left[ {\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}} = 1} \right) \wedge \left( {{1_{\left\{ {{{\tilde{D}}_j}} \right\}}} = 1} \right)|\bar{x}} \right] - {p_i}(\bar{x}){p_j}(\bar{x})} \right]} } . \\ \end{array} $$
(5.106)

5.1.4 Derivatives of the Decomposed Variance Terms

As both conditional variance terms are linear in the bivariate normal distribution, the derivative of the bivariate normal distribution will be calculated subsequently. Then, the derivatives of \( \begin{array} {c} b{V}\left[ {b{E}\left( {\tilde{L}|\{ {{\tilde{z}}_k}\} } \right)|\tilde{\bar{x}} = \bar{x}} \right] = \sum\limits_i {\sum\limits_j {{w_i}{w_j}ELG{D_i}ELG{D_j}} } \\ \cdot \left[ {{\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}(\bar{x})} \right),{\Phi^{ - 1}}\left( {{p_j}(\bar{x})} \right),\rho_{ij}^{\bar{x}}} \right) - {p_i}(\bar{x}){p_j}(\bar{x})} \right]. \\ \end{array} \) and \( \eta_{2,c}^\infty (\bar{x}) \) will be computed.

Proposition

The derivative of the bivariate normal distribution can be written as:

$$ \eta_{2,c}^{\text{GA}}(\bar{x}) $$
(5.107)

Proof

Using the notation

$$ \begin{array} {c} \frac{d}{{dx}}{\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right),{\Phi^{ - 1}}\left( {{p_j}\left( {\bar{x}} \right)} \right),\rho_{ij}^{\bar{x}}} \right) = \frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left( {{p_j}\left( {\bar{x}} \right)} \right) - \rho_{ij}^{\bar{x}} \cdot {\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right)}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) \\ + \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right) - \rho_{ij}^{\bar{x}} \cdot {\Phi^{ - 1}}\left( {{p_j}\left( {\bar{x}} \right)} \right)}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right). \\ \end{array} $$
(5.108)

and the chain rule, we get

$$ \begin{array} {c}{*{20}{c}} {{y_i}(\bar{x}) = \frac{{{\Phi^{ - 1}}(P{D_i}) - {c_i} \cdot \bar{x}}}{{\sqrt {{1 - {c_i}^2}} }}, \,\,\,{y_j}(\bar{x}) = \frac{{{\Phi^{ - 1}}(P{D_j}) - {c_j} \cdot \bar{x}}}{{\sqrt {{1 - {c_j}^2}} }}} \\ \end{array}, $$
(5.109)

For calculation of term (II) and (IV), we rewrite the bivariate normal distribution according to Appendix 2.8.6 as

$$ \begin{array} {c} \frac{d}{{d\bar{x}}}{\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right),{\Phi^{ - 1}}\left( {{p_j}\left( {\bar{x}} \right)} \right),\rho_{ij}^{\bar{x}}} \right) \\ = \frac{d}{{d\bar{x}}}{\Phi_2}\left( {{y_i}\left( {\bar{x}} \right),{y_j}\left( {\bar{x}} \right),\rho_{ij}^{\bar{x}}} \right) \\ = \underbrace {\frac{{d{y_i}}}{{d\bar{x}}}}_{(I)}\underbrace {\frac{\partial }{{\partial {y_i}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right)}_{(II)} + \underbrace {\frac{{d{y_j}}}{{d\bar{x}}}}_{(III)}\underbrace {\frac{\partial }{{\partial {y_j}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right)}_{(IV)}. \\ \end{array} $$
(5.110)

Thus, we have

$$ {\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right) = \int\limits_{z = - \infty }^{{y_j}} {\varphi (z)\Phi \left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}} \cdot z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)dz} . $$
(5.111)

The term (*) is equivalent to

$$ \begin{array} {c} \frac{\partial }{{\partial {y_i}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right) = \frac{\partial }{{\partial {y_i}}}\int\limits_{z = - \infty }^{{y_j}} {\varphi (z)\Phi \left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}} \cdot z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)dz} \\ = \frac{1}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}\int\limits_{z = - \infty }^{{y_j}} {\varphi (z)\varphi \left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}} \cdot z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)dz} \\ = \frac{1}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}\int\limits_{z = - \infty }^{{y_j}} {\frac{1}{{2\pi }}\exp \left( { - \frac{1}{2}\underbrace {\left[ {{z^2} + {{\left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}} \cdot z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)}^2}} \right]}_{{(*)}}} \right)dz} . \\ \end{array} $$
(5.112)

Hence, (5.111) can be written as

$$ \begin{array} {c} {z^2} + {\left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}} \cdot z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)^2} = \frac{{\left( {1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}} \right){z^2} + {y_i}^2 - 2{y_i}\rho_{ij}^{\bar{x}}z + \left( {\rho_{ij}^{\bar{x}}} \right){z^2}}}{{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} \\ = \frac{{{z^2} - 2{y_i}\rho_{ij}^{\bar{x}}z + {y_i}^2}}{{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} \\ = \frac{{{z^2} - 2{y_i}\rho_{ij}^{\bar{x}}z + {y_i}^2 + {y_i}^2{{\left( {\rho_{ij}^{\bar{x}}} \right)}^2} - {y_i}^2{{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}}{{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} \\ = {\left( {\frac{{z - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)^2} + {y_i}^2. \\ \end{array} $$
(5.113)

For solving the integral, we substitute \( \begin{array} {c} \frac{\partial }{{\partial {y_i}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right) = \frac{1}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}\int\limits_{z = - \infty }^{{y_i}} {\frac{1}{{2\pi }}\exp \left( { - \frac{1}{2}\left[ {{y_i}^2 + {{\left( {\frac{{z - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)}^2}} \right]} \right)dz} \\ = \varphi \left( {{y_i}} \right)\int\limits_{z = - \infty }^{{y_j}} {\frac{1}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}\varphi \left( {\frac{{z - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)dz} . \\ \end{array} \), and thus \( t: = \frac{{z - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{(\rho_{ij}^{\bar{x}})}^2}}} }} \). This leads to

$$ \frac{{dz}}{{dt}} = \sqrt {{1 - {{(\rho_{ij}^{\bar{x}})}^2}}} $$
(5.114)

Analogously, the term (IV) of (5.109) is equivalent to

$$ \begin{array} {c} \frac{\partial }{{\partial {y_i}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right) = \varphi \left( {{y_i}} \right)\int\limits_{t = - \infty }^{\frac{{{y_j} - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{(\rho_{ij}^{\bar{x}})}^2}}} }}} {\frac{1}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}\varphi (t)\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} dt} \\ = \varphi \left( {{y_i}} \right)\Phi \left( {\frac{{{y_j} - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right). \\ \end{array} $$
(5.115)

The derivatives (I) and (III) of (5.109) are given as

$$ \begin{array} {c} \frac{\partial }{{\partial {y_j}}}{\Phi_2}\left( {{y_i},{y_j},\rho_{ij}^{\bar{x}}} \right) = \frac{\partial }{{\partial {y_j}}}\int\limits_{z = - \infty }^{{y_i}} {(z)\Phi \left( {\frac{{{y_j} - \rho_{ij}^{\bar{x}}z}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right)dz} \\ = \varphi \left( {{y_j}} \right)\Phi \left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}}{y_j}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right). \\ \end{array} $$
(5.116)

Thus, inserting (5.114), (5.115), and (5.116) into (5.109), the derivative of the bivariate normal distribution finally results in

$$ \frac{{d{y_i}\left( {\bar{x}} \right)}}{{d\bar{x}}} = - \frac{{{c_i}}}{{\sqrt {{1 - {c_i}^2}} }} \,\,\,{\text{and}} \,\,\,\frac{{d{y_j}\left( {\bar{x}} \right)}}{{d\bar{x}}} = - \frac{{{c_j}}}{{\sqrt {{1 - {c_j}^2}} }}. $$
(5.117)

where the derivatives \( \begin{array} {c} \frac{d}{{d\bar{x}}}{\Phi_2}\left( {{y_i}\left( {\bar{x}} \right),{y_j}\left( {\bar{x}} \right),\rho_{ij}^{\bar{x}}} \right) = - \frac{{{c_i}}}{{\sqrt {{1 - {c_i}^2}} }}\varphi \left( {{y_i}} \right)\Phi \left( {\frac{{{y_j} - \rho_{ij}^{\bar{x}}{y_i}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) \\ - \frac{{{c_j}}}{{\sqrt {{1 - {c_j}^2}} }}\varphi \left( {{y_j}} \right)\Phi \left( {\frac{{{y_i} - \rho_{ij}^{\bar{x}}{y_j}}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) \\ = \frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_j}\left( {\bar{x}} \right)} \right] - \rho_{ij}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) \\ + \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}}\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right] - \rho_{ij}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_j}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right), \\ \end{array} \) and \( \frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}} \) are given by (5.16), which is equal to proposition (5.107).

As a next step, the derivatives of \( \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}} \) and \( \eta_{2,c}^\infty (\bar{x}) \) will be calculated. With

$$ \eta_{2,c}^{\text{GA}}(\bar{x}) $$
(5.118)

we get

$$ \begin{array} {c} \eta_{2,c}^\infty \left( {\bar{x}} \right) = \sum\limits_{i = 1}^n {\sum\limits_{j = 1}^n {{w_i}{w_j}ELG{D_i}ELG{D_j}} } \\ \cdot \left[ {{\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right),{\Phi^{ - 1}}\left( {{p_j}\left( {\bar{x}} \right)} \right),\rho_{ij}^{\bar{x}}} \right) - {p_i}\left( {\bar{x}} \right){p_j}\left( {\bar{x}} \right)} \right], \\ \end{array} $$
(5.119)

Using the derivative of the bivariate normal distribution from (5.117) yields

$$ \begin{array} {c} \frac{{d\eta_{2,c}^\infty \left( {\bar{x}} \right)}}{{d\bar{x}}} = \sum\limits_{i = 1}^n {\sum\limits_{j = 1}^n {{w_i}} } {w_j}ELG{D_i}ELG{D_j}\left[ {\frac{d}{{d\bar{x}}}{\Phi_2}\left( {{y_i}\left( {\bar{x}} \right),{y_j}\left( {\bar{x}} \right),\rho_{ij}^{\bar{x}}} \right) - \frac{d}{{d\bar{x}}}\left( {{p_i}\left( {\bar{x}} \right){p_j}\left( {\bar{x}} \right)} \right)} \right] \\ = \sum\limits_{i = 1}^n {\sum\limits_{j = 1}^n {{w_i}} } {w_j}ELG{D_i}ELG{D_j} \\ \cdot \left[ {\frac{d}{{d\bar{x}}}{\Phi_2}\left( {{y_i}\left( {\bar{x}} \right),{y_j}\left( {\bar{x}} \right),\rho_{ij}^{\bar{x}}} \right) - \left( {\frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}{p_j}\left( {\bar{x}} \right) + \frac{{d{p_j}\left( {\bar{x}} \right)}}{{d\bar{x}}}{p_i}\left( {\bar{x}} \right)} \right)} \right]. \\ \end{array} $$
(5.120)

Comparing the terms on the right-hand side, it can be found that the first and second summand as well as the third and fourth summand only differ concerning the indices i and j. Due to the double sum, each combination of i and j occurs twice. Thus, (5.120) can be simplified to:Footnote 65

$$ \rho_{ij}^{\bar{x}} = \rho_{ji}^{\bar{x}} $$
(5.121)

Similarly, the derivative of

$$ \begin{array} {c} \frac{{d\eta_{2,c}^\infty \left( {\bar{x}} \right)}}{{d\bar{x}}} = 2 \cdot \sum\limits_{i = 1}^n {\sum\limits_{j = 1}^n {{w_i}} } {w_j}ELG{D_i}ELG{D_j}\frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}} \\ \cdot \left( {\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_j}\left( {\bar{x}} \right)} \right] - \rho_{ij}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ij}^{\bar{x}}} \right)}^2}}} }}} \right) - {p_j}\left( {\bar{x}} \right)} \right). \\ \end{array} $$
(5.122)

is given as

$$ \begin{array} {lll} \eta_{2,c}^{\text{GA}}\left( {\bar{x}} \right) = \sum\limits_{i = 1}^n {{w_i}^2\left( {ELG{D_i}^2\left[ {{p_i}\left( {\bar{x}} \right) - {\Phi_2}\left( {{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right),{\Phi^{ - 1}}\left( {{p_i}\left( {\bar{x}} \right)} \right),\rho_{ii}^{\bar{x}}} \right)} \right]} \right.} \\ \left. { + VLG{D_i}{p_i}\left( {\bar{x}} \right)} \right) \\ \end{array} $$
(5.123)

Inserting the derivative of the bivariate normal distribution (5.117) finally leads to

$$ \frac{{d\eta_{2,c}^{\text{GA}}\left( {\bar{x}} \right)}}{{d\bar{x}}} = \sum\limits_{i = 1}^n {w_i^2} \left( {ELGD_i^2\left[ {\frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}} - \frac{d}{{d\bar{x}}}{\Phi_2}\left( {{y_i}\left( {\bar{x}} \right),{y_i}\left( {\bar{x}} \right),\rho_{ii}^{\bar{x}}} \right)} \right] + VLG{D_i}\frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}} \right). $$
(5.124)

5.1.5 Moment Matching in the BET-Model

5.1.5.1 Matching the First Moment

The expected loss of the original portfolio can be calculated as

$$\frac{{d\eta_{2,c}^{\text{GA}}\left( {\bar{x}} \right)}}{{d\bar{x}}} = \sum\limits_{i = 1}^n {w_i^2} \frac{{d{p_i}\left( {\bar{x}} \right)}}{{d\bar{x}}}\cdot \left( {ELGD_i^2\left[ {1 - 2\Phi \left( {\frac{{{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right] - \rho_{ii}^{\bar{x}}{\Phi^{ - 1}}\left[ {{p_i}\left( {\bar{x}} \right)} \right]}}{{\sqrt {{1 - {{\left( {\rho_{ii}^{\bar{x}}} \right)}^2}}} }}} \right)} \right]} \right. \left. { + VLG{D_i}} \right).$$
(5.125)

and the expected loss of the hypothetical portfolio as

$$ b{E}\left( {{{\tilde{L}}^{\text{orig}}}} \right) = \sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot LGD \cdot b{E}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} \right)} } = \sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot LGD \cdot P{D_{s,i}}} } $$
(5.126)

with \( \begin{array} {c} b{E}\left( {{{\tilde{L}}^{\text{hyp}}}} \right) = \sum\limits_{i = 1}^D {\frac{1}{D} \cdot LGD \cdot b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} \right)} = \frac{1}{D} \cdot LGD \cdot \sum\limits_{i = 1}^D {\bar{p}} \\ = \frac{1}{D} \cdot LGD \cdot D \cdot \bar{p} = LGD \cdot \bar{p}, \\ \end{array} \) for all i. Thus, matching the expectation for both portfolios leads to

$$ b{E}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} \right) = \bar{p} $$
(5.127)

5.1.5.2 Matching the Second Moment

For the original portfolio, the variance can be calculated as

$$ \begin{array} {lll} & b{E}\left( {{{\tilde{L}}^{\text{orig}}}} \right)\mathop { = }\limits^! b{E}\left( {{{\tilde{L}}^{\text{hyp}}}} \right) \\ \Leftrightarrow & \sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot LGD \cdot P{D_{s,i}}} } = LGD \cdot \bar{p} \\ \Leftrightarrow & \bar{p} = \sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot P{D_{s,i}}} } . \\ \end{array} $$
(5.128)

As the default variable is Bernoulli distributed, the variance terms equal

$$ \begin{array} {c} b{V}\left( {{{\tilde{L}}^{\text{orig}}}} \right) = b{V}\left( {\sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot LGD \cdot {1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} } } \right) \\ = LG{D^2} \cdot b{V}\left( {\sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot {1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} } } \right) \\ = LG{D^2} \cdot {\text{Cov}}\left( {\sum\limits_{s = 1}^S {\sum\limits_{i = 1}^{{n_s}} {{w_{s,i}} \cdot {1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} }, \sum\limits_{t = 1}^S {\sum\limits_{j = 1}^{{n_t}} {{w_{t,j}} \cdot {1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} } } \right) \\ = LG{D^2} \cdot \sum\limits_{s = 1}^S {\sum\limits_{t = 1}^S {\sum\limits_{i = 1}^{{n_s}} {\sum\limits_{j = 1}^{{n_t}} {{w_{s,i}} \cdot {w_{t,j}} \cdot {\text{Cov}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)} } } } \\ = LG{D^2} \cdot \sum\limits_{s = 1}^S {\sum\limits_{t = 1}^S {\sum\limits_{i = 1}^{{n_s}} {\sum\limits_{j = 1}^{{n_t}} {{w_{s,i}} \cdot {w_{t,j}} \cdot {\text{Corr}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) \cdot \sqrt {{b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} \right)}} \cdot \sqrt {{b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)}} .} } } } \\ \end{array} $$
(5.129)

and we obtain

$$ b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} \right) = P{D_{s,i}} \cdot \left( {1 - P{D_{s,i}}} \right)\quad {\text{and}}\quad b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) = P{D_{t,j}} \cdot \left( {1 - P{D_{t,j}}} \right) $$
(5.130)

Due to the independence of the default events in the hypothetical portfolio, the variance of this portfolio is

$$ \begin{array} {c} b{V}\left( {{{\tilde{L}}^{\text{orig}}}} \right) = LG{D^2} \cdot \sum\limits_{s = 1}^S {\sum\limits_{t = 1}^S {\sum\limits_{i = 1}^{{n_s}} {\sum\limits_{j = 1}^{{n_t}} {{w_{s,i}} \cdot {w_{t,j}} \cdot {\text{Corr}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)} } } } \\ \cdot \sqrt {{P{D_{s,i}}\left( {1 - P{D_{s,i}}} \right)P{D_{t,j}}\left( {1 - P{D_{t,j}}} \right)}} . \\ \end{array} $$
(5.131)

Matching the variance terms (5.130) and (5.131) leads to

$$ \begin{array} {c} b{V}\left( {{{\tilde{L}}^{\text{hyp}}}} \right) = b{V}\left( {\sum\limits_{i = 1}^D {\frac{1}{D} \cdot LGD \cdot {1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} } \right) = \frac{1}{{{D^2}}} \cdot LG{D^2} \cdot b{V}\left( {\sum\limits_{i = 1}^D {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} } \right) \\ = \frac{1}{{{D^2}}} \cdot LG{D^2} \cdot D \cdot b{V}\left( {{1_{\left\{ {{{\tilde{D}}_i}} \right\}}}} \right) = \frac{1}{D} \cdot LG{D^2} \cdot \bar{p} \cdot \left( {1 - \bar{p}} \right). \\ \end{array} $$
(5.132)

5.1.6 Interrelation of the Pairwise Default Correlation and the Asset Correlation

Using the standard calculus for the correlation and covariance as well as the variance of a Bernoulli distributed variable, the pairwise default correlation between borrower i in sector s and borrower j in sector t can be expressed as

$$ \begin{array} {lll} & b{V}\left( {{{\tilde{L}}^{\text{orig}}}} \right)\mathop { = }\limits^! b{V}\left( {{{\tilde{L}}^{\text{hyp}}}} \right) \\ \Leftrightarrow & D = \frac{{\bar{p} \cdot \left( {1 - \bar{p}} \right)}}{{\sum\limits_{s = 1}^S {\sum\limits_{t = 1}^S {\sum\limits_{i = 1}^{{n_s}} {\sum\limits_{j = 1}^{{n_t}} {{w_{s,i}} \cdot {w_{t,j}} \cdot {\text{Corr}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)} } } } \cdot \sqrt {{P{D_{s,i}}\left( {1 - P{D_{s,i}}} \right)P{D_{t,j}}\left( {1 - P{D_{t,j}}} \right)}} }}. \\ \end{array} $$
(5.133)

The expectation values of the individual default events equal \( \begin{array} {c} {\text{Corr}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) = \frac{{{\text{Cov}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)}}{{\sqrt {{b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} \right)}} \cdot \sqrt {{b{V}\left( {{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)}} }} \\ = \frac{{{\text{Cov}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)}}{{\sqrt {{P{D_{s,i}} \cdot \left( {1 - P{D_{s,i}}} \right) \cdot P{D_{t,j}} \cdot \left( {1 - P{D_{t,j}}} \right)}} }} \\ = \frac{{b{E}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}} \cdot {1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) - b{E}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}}} \right) \cdot b{E}\left( {{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right)}}{{\sqrt {{P{D_{s,i}} \cdot \left( {1 - P{D_{s,i}}} \right) \cdot P{D_{t,j}} \cdot \left( {1 - P{D_{t,j}}} \right)}} }}. \\ \end{array} \) and \( P{D_{s,i}} \). Similar to (5.102), assuming a normally distributed asset return, the expectation value of a simultaneous default can be written as

$$ P{D_{t,j}} $$
(5.134)

Thus, we get

$$ \begin{array} {c} b{E}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}} \cdot {1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) = b{P}\left[ {\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}} = 1} \right) \wedge \left( {{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}} = 1} \right)} \right] \\ = b{P}\left( {{{\tilde{a}}_{s,i}} \leq {\Phi^{ - 1}}(P{D_{s,i}}),{{\tilde{a}}_{t,j}} \leq {\Phi^{ - 1}}(P{D_{t,j}})} \right) \\ = {\Phi_2}\left( {{\Phi^{ - 1}}\left( {P{D_{s,i}}} \right),{\Phi^{ - 1}}\left( {P{D_{t,j}}} \right),{\text{Corr}}\left( {{{\tilde{a}}_{s,i}},{{\tilde{a}}_{t,j}}} \right)} \right). \\ \end{array} $$
(5.135)

5.1.7 Expected Number of Defaults in the Infectious Defaults Model

Due to the homogeneity of the portfolio and the stochastic independence of all indicator variables, the expected number of defaults is

$$ {\text{Corr}}\left( {{1_{\left\{ {{{\tilde{D}}_{s,i}}} \right\}}},{1_{\left\{ {{{\tilde{D}}_{t,j}}} \right\}}}} \right) = \frac{{{\Phi_2}\left( {{\Phi^{ - 1}}\left( {P{D_{s,i}}} \right),{\Phi^{ - 1}}\left( {P{D_{t,j}}} \right),{\text{Corr}}\left( {{{\tilde{a}}_{s,i}},{{\tilde{a}}_{t,j}}} \right)} \right) - P{D_{s,i}} \cdot P{D_{t,j}}}}{{\sqrt {{P{D_{s,i}}\left( {1 - P{D_{s,i}}} \right)P{D_{t,j}}\left( {1 - P{D_{t,j}}} \right)}} }} $$
(5.136)

Rights and permissions

Reprints and permissions

Copyright information

© 2010 Physica-Verlag HD

About this chapter

Cite this chapter

Hibbeln, M. (2010). Model-Based Measurement of Sector Concentration Risk in Credit Portfolios. In: Risk Management in Credit Portfolios. Contributions to Economics. Physica, Heidelberg. https://doi.org/10.1007/978-3-7908-2607-4_5

Download citation

Publish with us

Policies and ethics