Skip to main content
Log in

Nearest neighbor hazard estimation with left-truncated duration data

  • Original Paper
  • Published:
AStA Advances in Statistical Analysis Aims and scope Submit manuscript

Abstract

Duration data often suffer from both left-truncation and right-censoring. We show how both deficiencies can be overcome at the same time when estimating the hazard rate nonparametrically by kernel smoothing with the nearest-neighbor bandwidth. Smoothing Turnbull’s estimator of the cumulative hazard rate, we derive strong uniform consistency of the estimate from Hoeffding’s inequality, applied to a generalized empirical distribution function. We also apply our estimator to rating transitions of corporate loans in Germany.

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

Similar content being viewed by others

References

  • Bluhm, C., Overbeck, L., Wagner, C.: An Introduction to Credit Risk Modeling. Chapman & Hall, London (2002)

    Book  Google Scholar 

  • Einmahl, U., Mason, D.: Uniform in bandwidth consistency of kernel-type function estimators. Ann. Stat. 33, 1380–1403 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  • Gefeller, O., Dette, H.: Nearest neighbour kernel estimation of the hazard function from censored data. J. Stat. Comput. Simul. 43, 93–101 (1992)

    Article  Google Scholar 

  • Gefeller, O., Weißbach, R., Bregenzer, T.: The implementation of a data-driven selection procedure for the smoothing parameter in nonparametric hazard rate estimation using SAS/IML software. In: Friedl, H., Berghold, A., Kauermann, G. (eds.) Proceedings of the 13th SAS European Users Group International Conference, Stockholm, pp. 1288–1300. SAS Institute, Carry (1996)

    Google Scholar 

  • Goto, F.: Achieving semiparametric efficiency bounds in left-censored duration models. Econometrica 64(2), 439–442 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  • Grillenzoni, C.: Robust nonparametric estimation of the intensity function of point data. AStA Adv. Stat. Anal. 92, 117–134 (2008)

    Article  MathSciNet  Google Scholar 

  • Hewitt, E., Savage, L.: Symmetric measures on Cartesian products. Trans. Am. Math. Soc. 80, 470–501 (1955)

    Article  MathSciNet  MATH  Google Scholar 

  • Hoeffding, W.: Probability inequalities for sums of bounded random variables. J. Am. Stat. Assoc. 58, 13–30 (1963)

    Article  MathSciNet  MATH  Google Scholar 

  • Kiefer, N., Larson, C.: A simulation estimator for testing the time homogeneity of credit rating transitions. J. Empir. Finance 14, 818–835 (2007)

    Article  Google Scholar 

  • Kim, Y.-D., James, L., Weißbach, R.: Bayesian analysis of multi-state event history data: Beta–Dirichlet process prior. Biometrika 99, 127–140 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Knüppel, L., Hermsen, O.: Median split, k-group split, and optimality in continuous populations. AStA Adv. Stat. Anal. 94, 53–74 (2010)

    Article  MathSciNet  Google Scholar 

  • Li, D., Li, Q.: Nonparametric/semiparametric estimation and testing of econometric models with data dependent smoothing parameters. J. Econom. 157, 179–190 (2010)

    Article  Google Scholar 

  • Merton, R.: On the pricing of corporate debt: the risk structure of interest rates. J. Finance 29, 449–470 (1974)

    Google Scholar 

  • Schäfer, H.: Local convergence of empirical measures in the random censorship situation with application to density and rate estimators. Ann. Stat. 14, 1240–1245 (1986)

    Article  MATH  Google Scholar 

  • Shorack, G., Wellner, J.: Empirical Processes with Application to Statistics. Wiley, New York (1986)

    Google Scholar 

  • Silverman, B.: Density Estimation. Chapman & Hall, London (1986)

    MATH  Google Scholar 

  • Stute, W.: Almost sure representations of the product-limit estimator for truncated data. Ann. Stat. 21, 146–156 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  • Turnbull, B.W.: The empirical distribution function with arbitrarily grouped, censored and truncated data. J. R. Stat. Soc., Ser. B, Stat. Methodol. 38, 290–295 (1976)

    MathSciNet  MATH  Google Scholar 

  • Weißbach, R., Mollenhauer, T.: Modelling rating transitions. J. Korean Stat. Soc. 4, 469–485 (2011)

    Article  Google Scholar 

  • Weißbach, R., Pfahlberg, A., Gefeller, O.: Double-smoothing in kernel hazard rate estimation. Methods Inf. Med. 47, 167–173 (2008)

    Google Scholar 

  • Weißbach, R., Tschiersch, P., Lawrenz, C.: Testing time-homogeneity of rating transitions after origination of debt. Empir. Econ. 36, 575–596 (2009)

    Article  Google Scholar 

  • Weißbach, R., Walter, R.: A likelihood ratio test for stationarity of rating transitions. J. Econom. 155, 188–194 (2010)

    Article  Google Scholar 

  • Weißbach, R.: A general kernel functional estimator with general bandwidth—strong consistency and applications. J. Nonparametr. Stat. 18, 1–12 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  • Wied, D., Weißbach, R.: Consistency of the kernel density estimator—a survey. Stat. Pap. 53, 1–21 (2012)

    Article  MATH  Google Scholar 

  • Woodroofe, M.: Estimating a distribution function with truncated data. Ann. Stat. 13, 163–177 (1985)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgement

Financial support by Deutsche Forschungsgemeinschaft is gratefully acknowledged (SFB 823 and Grant WE3573/2).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rafael Weißbach.

Appendices

Appendix A: Proof of Theorem 2

The proof of Theorem 2 is in four steps. First, for an interval I:=[a,b]⊆[A,B] we establish an exponential bound for the distribution of the difference \(|\varPsi _{n}^{*}(I) - \varPsi(I)|\):

(9)

for all p>0, ε>0, n∈ℕ>0 and for each fixed I⊆[A,B] with Ψ(I)≤p.

Because of definition (4) and the boundedness of \(0 \le \varDelta ^{x}_{i} \le \varDelta _{\max} < \infty\)

(10)

is the arithmetic mean of the n independent and bounded random variables for each fixed I⊆[A,B], distributed as

The expectation, the variance and the bound of T I can then be calculated for fixed I⊆[A,B] with Ψ(I)≤p.

The expectation of T I follows from assumption (B3):

(11)

From assumption (B1), we get the following bound of |T I | on [A,B]:

(12)

The variance of T I can be obtained from the expectation (11) and the bound (12) as follows:

(13)

From (10), (11), (12), (13), and the inequality from Hoeffding (1963) results the following right bound:

for each fixed interval I⊆[A,B] with Ψ(I)≤p.

In the second step we derive the inequality

(14)

almost surely for a constant \(C>\sqrt{2(2\varDelta _{\max} M + \varPsi(B))}\) and large n.

On the right hand side of the inequality (9), p and ε can be substituted with p n and \(\varepsilon_{n}:=C\sqrt{\log{(n)}p_{n}/n}\) for C>0 and n>1 altering the upper bound to

The series (A n ) is then summable starting from some large n<∞ only if the exponent β n :=(C 2 p n )/(2g(p n +ε n ))>1. From \(\varepsilon_{n}/p_{n} = C\sqrt{\log(n)/(np_{n})}\) and the assumptions for p n follow ε n /p n →0 and p n /(p n +ε n )→1 for large n. The condition β n >1 can be then achieved with C 2/2g>1 or \(C > \sqrt{2g}\).

As a consequence, the series (A n ) is summable from some large n<∞ and only for \(C>\sqrt{2g}\). For each I⊆[A,B] with Ψ(I)≤p n we get then \(\exists C>\sqrt{2g}\)m<∞, \(m \in \mathbb{N}: \sum_{n=m}^{\infty} P(|\varPsi _{n}^{*}(I)-\varPsi(I)| > \varepsilon_{n}) < \sum_{n=m}^{\infty} A_{n} < \infty\) and \(\forall m < \infty, m \in \mathbb{N}: \sum_{n=1}^{m} P(|\varPsi _{n}^{*}(I)-\varPsi(I)| > \varepsilon_{n}) \le m < \infty\).

Because of the summability of \(P(|\varPsi _{n}^{*}(I)-\varPsi(I)| > \varepsilon_{n})\),

$$P\Bigl( \limsup_{n \rightarrow \infty} \big|\varPsi _n^*(I)-\varPsi(I)\big| > \varepsilon_n \Bigr)=0 $$

results from the Borel–Cantelli lemma for \(C>\sqrt{2g}\), i.e. \(|\varPsi _{n}^{*}(I)-\varPsi(I)|\) does not exceed ε n for most of the n. For large n and for all I⊆[A,B] with Ψ(I)≤p n , we derive almost surely that \(|\varPsi _{n}^{*}(I)-\varPsi(I)| \le C\sqrt{\log{(n)}p_{n}/n}\).

The same inequality holds for the supremum of \(|\varPsi _{n}^{*}(I)-\varPsi(I)|\) on [A,B]: \(\sup_{I \subseteq [A,B], \varPsi(I)\le p_{n}}|\varPsi _{n}^{*}(I)-\varPsi(I)| \le C\sqrt{\log{(n)}p_{n}/n}\) for \(C>\sqrt{2g}\) and large n almost surely.

Using the results above we prove the following inequality in a third step:

$$\sup_{I \subseteq [A,B], \varPsi(I)\le p_n}\big|\varPsi _n^*(I)-\varPsi (I)\big| \le C\cdot p_n \sqrt {\log {(n)} / n} $$

almost surely for some C>D G M and large n.

From assumption (B4) and the limes superior formulation of Hewitt and Savage (1955) we obtain the right bound

$$G_n(x)- G(x) \le \big|G_n(x)- G(x)\big| \le \sup_{x \in [A,B]}\big|G_n(x)- G(x)\big| \le C_1'\cdot\sqrt {\log {(n)} / n} $$

almost surely for \(C_{1}'>D_{G}\), large n and all x∈[A,B]. These bounds can be rewritten for G n (x) as follows:

$$G_n(x) \ge G(x) - C_1'\sqrt {\log {(n)} / n} \ge \inf_{t \in [A,B]}G(t) - C_1'\cdot\sqrt {\log {(n)} / n}. $$

From assumption (B4) we have inf t∈[A,B] G(t)>0. Because of \(\sqrt {\log {(n)} / n} \rightarrow 0\), the following inequalities hold for x∈[A,B] and large n:

and

The following bounds for \(\varPsi _{n}^{*}(I)-\varPsi(I)\) and \(\varPsi _{n}^{*}(I)\) result from (14) almost surely for I⊆[A,B] with Ψ(I)≤p n , large n and \(C_{2}'>\sqrt{2\cdot(2\varDelta _{\max} M + \varPsi(B))}\):

and consequently

$$\varPsi _n^*(I) \le \varPsi(I) + C_2'\sqrt{\log{(n)}p_n/n} \le p_n + C_2'\sqrt{\log{(n)}p_n/n}. $$

We then obtain the following equation from assumption (B2) almost surely for each I⊆[A,B] with Ψ(I)≤p n and large n:

By \(p_{n} + C_{2}'\sqrt{\log{(n)}p_{n}/n} = p_{n}[1 + C_{2}'\sqrt{\log{(n)}/(p_{n}n)}]\) it is evident that \(C_{2}'\sqrt{\log{(n)}/(p_{n}n)}\) can be neglected for large n because of the assumptions for p n . For large n, we can also neglect the term \(\sqrt {\log {(n)} / n}\) in the numerator. For all I⊆[A,B] with Ψ(I)≤p n and for large n, we derive the inequality

$$\big|\varPsi _n^*(I)-\varPsi (I)\big| \le \frac{C_1'}{\inf_{t \in [A,B]}G(t)}p_n\sqrt {\log {(n)} / n}=C_1'\cdot M \cdot p_n\sqrt {\log {(n)} / n} $$

almost surely.

The requested bound

$$\sup_{I \subseteq [A,B], \varPsi(I)\le p_n}\big|\varPsi _n^*(I)-\varPsi (I)\big| \le C\cdot p_n\sqrt {\log {(n)} / n} $$

results for some C>D G M and large n almost surely.

In a final step we examine the expression \(\sup_{I \subseteq [A,B], \varPsi(I)\le p_{n}}|\varPsi _{n}(I)-\varPsi (I)|\). This overall difference can be represented by the sum of the deviations of the empirical and theoretical measures Ψ n (I) and Ψ(I) from the preliminary measure \(\varPsi _{n}^{*}( I)\) as follows:

Because \(p_{n}\sqrt{\log (n)/n}/\sqrt{\log (n)p_{n}/n} = \sqrt{p_{n}}\) approaches zero, i.e. \(p_{n}\sqrt{\log (n)/n} \le \sqrt{\log (n)p_{n}/n}\) holds for large n.

The previously mentioned upper bounds of \(|\varPsi _{n}(I)-\varPsi _{n}^{*}(I)|\) and \(|\varPsi _{n}^{*}(I)-\varPsi (I)|\) imply the existence of a constant \(C > \sqrt{2\cdot(2\varDelta _{\max} M + \varPsi(B))} + D_{G} \cdot M\), such that almost surely for large n

Due to the symmetry of Ψ n (I) the limes superior formulation of the convergence follows from Hewitt and Savage (1955).

Appendix B: Proof of Corollary 3

The boundedness of the \(\varDelta ^{x}_{i}\) for each x∈[A,B] and conditions (B1) and (B2) follow from the definition of \(\varDelta ^{x}_{i}\). This is so because the variables \(\varDelta ^{x}_{i}\) do not depend on the x.

The consistency of the estimator G n (⋅) (B4) can be easily shown and is a slight modification of the law of the iterated logarithm (see Shorack and Wellner 1986, p. 504).

The assumption (A2) for F(⋅) implies that the cumulative hazard rate Λ(⋅) is strictly increasing and the hazard rate λ(⋅) is obviously strictly positive on [A,B].

Now only the condition (B3) needs to be verified. We note that the vectors S i =(X i ,L i ,δ i ) i=1,…,n are observable under L i X i . Hence, we derive the following conditional expectation:

(15)

where F X,δ(x,y)=P(Xx,δyLX) is the conditional distribution function of (X,δ).

The integral \(\int_{x_{1} \in I} dF^{X, \delta}(x_{1},1)\) for the intervals I:=[a,b]⊆[A,B] can now be calculated. First we express the probability P(X i I,δ i =1∣L i X i ) in the terms of the non-observable vector (T i ,L i ,C i ) as follows:

(16)

where α=P(L i X i ). Hence, we express the probabilities P(X i I,δ i =1∣L i X i ) and P(T i I,L i T i C i ) as the following expectations of the Bernoulli-variables:

(17)

and

(18)

One can see that dF X,δ(x,1)=α −1 F L(j)(1−F C(j)) dF(x) follows from the expressions (16), (17), and (18). Consequently the expectation (15) can be written as follows:

Obviously, conditions (B1)–(B4) hold and the local convergence

follows for a constant \(D \le 2(\sqrt{2\cdot(2M + \varLambda(B))} + 2 M)\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Weißbach, R., Poniatowski, W. & Krämer, W. Nearest neighbor hazard estimation with left-truncated duration data. AStA Adv Stat Anal 97, 33–47 (2013). https://doi.org/10.1007/s10182-012-0194-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10182-012-0194-5

Keywords

Navigation