Quantile Estimation When Applying Conditional Monte Carlo
Marvin K. Nakayama
Computer Science Department, New Jersey Institute of Technology, Newark, NJ 07102, U.S.A.
Keywords:
Quantile, Value-at-Risk, Variance Reduction, Conditional Monte Carlo, Confidence Interval.
Abstract:
We describe how to use conditional Monte Carlo (CMC) to estimate a quantile. CMC is a variance-reduction
technique that reduces variance by analytically integrating out some of the variability. We show that the
CMC quantile estimator satisfies a central limit theorem and Bahadur representation. We also develop three
asymptotically valid confidence intervals (CIs) for a quantile. One CI is based on a finite-difference estima-
tor, another uses batching, and the third applies sectioning. We present numerical results demonstrating the
effectiveness of CMC.
1 INTRODUCTION
For a continuous random variable X with a strictly
increasing cumulative distribution function (CDF) F
and fixed 0 < p < 1, the p-quantile of X is defined as
the constant ξ such that P(X ξ) = p. A well-known
example is the median, which is the 0.5-quantile. The
p-quantile ξ also can equivalently be expressed as ξ =
F
1
(p).
Quantiles are often used in application areas to
measure risk. For example, in finance, a quantile is
known as a value-at-risk, and quantiles are widely em-
ployed to assess portfolio risk. For example, bank-
ing regulations specify capital requirements for a firm
in terms of 0.99-quantiles of the random loss (Jorion,
2007). In nuclear engineering, safety and uncertainty
analyses are often performed with a 0.95-quantile
(U.S. Nuclear Regulatory Commission, 1989).
Suppose that we have a simulation model that out-
puts a random variable X. When applying simple ran-
dom sampling (SRS), the typical approach to estimate
the p-quantile ξ is to run independent and identically
distributed (i.i.d.) replications of the model, and form
an estimator of the CDF from the sample outputs. In-
verting the CDF estimator yields a quantile estimator.
Because of the noise inherent in any stochastic
simulation, the quantile estimator has some error,
which should be measured. A standard way of as-
sessing the error is by forming a confidence interval
for the true quantile ξ. For example, the U.S. Nu-
clear Regulatory Commission requires nuclear plant
licensees to satisfy a so-called 95/95 criterion, which
entails establishing, with 95% confidence, that the
0.95-quantile lies below a mandated threshold; see
Section 24.9 of (U.S. Nuclear Regulatory Commis-
sion, 2011). Thus, we need not only a point estimate
of a quantile but also a confidence interval for it.
There are several approaches to construct a CI
when applying SRS. One technique, which is some-
times called the nonparametric method, exploits a bi-
nomial property of the i.i.d. sample; see Section 2.6.1
of (Serfling, 1980). Another way first shows that
the quantile estimator satisfies a central limit theorem
(CLT), and then unfolds the CLT to obtain a CI. The
key to applying this technique is consistently estimat-
ing the asymptotic variance constant appearing in the
CLT; approaches for accomplishing this include using
a finite difference (Bloch and Gastwirth, 1968; Bofin-
ger, 1975) and kernel methods (Falk, 1986). Rather
than consistently estimating the asymptotic variance,
we can instead apply batching or sectioning, the latter
of which was originally developed for SRS in Sec-
tion III.5a of (Asmussen and Glynn, 2007) and ex-
tended in (Nakayama, 2014a) to work when applying
the variance-reductiontechniques control variates and
importance sampling. Batching and sectioning divide
the i.i.d. outputs into independent batches, computing
a quantile estimator from each batch, and construct-
ing a CI from the batch quantile estimators.
In this paper, we use conditional Monte Carlo to
estimate a quantile. CMC reduces variance (com-
pared to SRS) by analytically integrating out the vari-
ability that remains after conditioning on an auxiliary
random variable Y; e.g., see Section 8.3 of (Ross,
2006) or Section V.4 of (Asmussen and Glynn, 2007).
We prove that the CMC quantile estimator satisfies a
280
K. Nakayama M..
Quantile Estimation When Applying Conditional Monte Carlo.
DOI: 10.5220/0005109702800285
In Proceedings of the 4th International Conference on Simulation and Modeling Methodologies, Technologies and Applications (SIMULTECH-2014),
pages 280-285
ISBN: 978-989-758-038-3
Copyright
c
2014 SCITEPRESS (Science and Technology Publications, Lda.)
CLT and a Bahadur representation (Bahadur, 1966).
The latter shows that a quantile estimator can be ap-
proximated as the true quantile plus a linear transfor-
mation of the corresponding CDF estimator, with a
remainder term that vanishes at some rate as the sam-
ple size grows. Since the CDF estimator is typically
a sample average, it satisfies a CLT under appropri-
ate conditions. Thus, the Bahadur representation pro-
vides insight into why a quantile estimator, which is
not a sample average, satisfies a CLT. It also allows us
to construct asymptotically valid CIs for ξ by using a
finite difference or sectioning, and we develop those
CIs in this paper.
CMC has previously been employed to derive an
estimator of a sensitivity of a quantile with respect to
a model parameter (Fu et al., 2009). For example,
suppose a financial investor has a portfolio of loans,
each of which may default. The investor may want to
estimate the sensitivity of the 0.99-quantile of the loss
of the portfolio, where the sensitivity is taken with
respect to a parameter of the loss distribution of an
individual obligor. While (Fu et al., 2009) apply CMC
to estimate quantile sensitivities, the method has not
been used (to the best of our knowledge) to estimate
the quantile itself.
The rest of the paper develops as follows. Sec-
tion 2 reviews how to apply SRS to estimate and con-
struct CIs for a quantile ξ. Section 3 develops our
CMC estimator of a quantile, shows that it satisfies a
CLT and Bahadur representation, and uses these re-
sults to construct CIs for ξ. Section 4 presents nu-
merical results from a simple model, and we provide
concluding remarks in Section 5. Proofs of the results
are given in (Nakayama, 2014b).
2 SIMPLE RANDOM SAMPLING
Let X be a random variable with CDF F. We first re-
view how to estimate and construct confidence inter-
vals for the p-quantile ξ = F
1
(p) inf{x : F(x)
p} of F (or equivalently of X) for a fixed 0 < p < 1
when applying simple random sampling (SRS).
Let X
i
, i = 1, 2,.. .,n, be a sample of n i.i.d. ob-
servations from F. The SRS estimator of F(x) =
E[I(X x)] is the empirical distribution function
F
n
(x) = (1/n)
n
i=1
I(X
i
x). The SRS p-quantile es-
timator is ξ
n
= F
1
n
(p). We can alternatively compute
ξ
n
by first sorting the sample X
1
,X
2
,... ,X
n
into the or-
der statistics X
(1)
X
(2)
··· X
(n)
, and then setting
ξ
n
= X
(np)
, where · denotes the ceiling function.
Section 2.3 of (Serfling, 1980) provides an
overview of ξ
n
and its properties. For example, let
f denote the derivative (when it exists) of F. If
f(ξ) > 0, then ξ
n
satisfies the following CLT:
n(ξ
n
ξ) N(0, p(1p)/ f
2
(ξ)) (1)
as n , where denotes convergence in distri-
bution (e.g., see Chapter 5 of (Billingsley, 1995)),
and N(a,b
2
) is a normal random variable with mean
a and variance b
2
. Moreover, ξ
n
also satisfies a
so-called (weak) Bahadur representation (Bahadur,
1966; Ghosh, 1971):
ξ
n
= ξ
F
n
(ξ) p
f(ξ)
+ R
n
, with
nR
n
0 (2)
as n . By (2), the left side of (1) equals
n(F
n
(ξ) p)/ f(ξ) +
nR
n
, where the first term
converges weakly to the right side of (1), and the sec-
ond weakly vanishes by (2). Thus, the Bahadur repre-
sentation provides insight into why ξ
n
, which is not a
sample average, satisfies a CLT, as it can be approxi-
mated in terms of the empirical distribution, which is
a sample mean.
(Ghosh, 1971) also establishes a version of (2) for
the p
n
-quantile with perturbed p
n
that converges to
p, rather than the p-quantile for fixed p. This vari-
ation can be useful for constructing a consistent es-
timator of λ 1/ f(ξ), which appears in the asymp-
totic variance in (1) and can be used to construct a
confidence interval for ξ. If f(ξ) > 0, then for any
p
n
= p+ O(n
1/2
), the SRS estimator F
1
n
(p
n
) of the
p
n
-quantile F
1
(p
n
) satisfies
F
1
n
(p
n
) = ξ
p
n
F
n
(ξ) p
f(ξ)
+R
n
, with
nR
n
0
(3)
as n , where
ξ
p
n
= ξ+ (p
n
p)/ f(ξ). (4)
To see how to use these results to consis-
tently estimate λ, first note that λ =
d
dp
F
1
(p) =
lim
h0
[F
1
(p+ h) F
1
(ph)]/2h. This suggests
estimating λ with the finite difference
λ
n
=
F
1
n
(p+ h
n
) F
1
n
(ph
n
)
2h
n
, (5)
where h
n
> 0 is known as the bandwidth. The terms in
the numerator of the finite difference are precisely in
the form of (3) with p
n
= p±h
n
, which allows prov-
ing λ
n
λ as n when h
n
= cn
1/2
for any con-
stant c > 0; see Section 2.6.3 of (Serfling, 1980). Us-
ing a different proof technique, (Bloch and Gastwirth,
1968) and (Bofinger, 1975) also show λ
n
λ when
f is continuous in a neighborhood of ξ, and h
n
0
and nh
n
as n . Thus, unfolding the CLT
in (1) leads to the following two-sided (1 α)-level
(0 < α < 1) confidence interval for ξ:
I
n
= [ξ
n
±z
α
p
p(1 p)λ
n
/
n], (6)
QuantileEstimationWhenApplyingConditionalMonteCarlo
281
where z
α
= Φ
1
(1α/2) and Φ is the CDF of a stan-
dard (mean 0 and unit variance) normal. The CI I
n
is asymptotically valid in the sense that P(ξ I
n
)
1α as n .
Determining an appropriate value for the band-
width h
n
in the finite difference can be difficult in
practice. Alternatively, we can avoid trying to con-
sistently estimate λ by instead applying batching or
sectioning. In batching, we divide the n outputs
X
1
,X
2
,... ,X
n
into b 2 equal-sized batches, where
the jth batch, j = 1,2,.. .,b, consists of the m =
n/b outputs X
( j1)m+i
, i = 1,2,. .. ,m. A reasonable
choice for the number of batches is b = 10. For
each batch j, we define the CDF estimator F
j,m
(x) =
(1/m)
m
i=1
I(X
( j1)m+i
x) and corresponding p-
quantile estimator ξ
j,m
= F
1
j,m
(p). Since the n out-
puts are i.i.d., the b batches are i.i.d., so ξ
j,m
, j =
1,2, . ..,b, are i.i.d. We compute their sample aver-
age
¯
ξ
b,m
= (1/b)
b
j=1
ξ
j,m
and their sample variance
S
2
b,m
= (1/(b1))
b
j=1
(ξ
j,m
¯
ξ
b,m
)
2
. An asymptot-
ically valid (as m with b 2 fixed) (1α)-level
CI for ξ using batching is then
J
n
= [
¯
ξ
b,m
±t
α
S
b,m
/
b],
where t
α
= T
1
b1
(1 α/2) and T
b1
is the CDF of a
Student t distribution with b1 degrees of freedom.
Similar to batching, sectioning was originally de-
veloped in Section III.5a of (Asmussen and Glynn,
2007) for SRS, and it replaces the batching point es-
timator
¯
ξ
b,m
with the overall quantile estimator ξ
n
.
Specifically, let S
2
b,m
= (1/(b 1))
b
j=1
(ξ
j,m
ξ
n
)
2
,
and the sectioning two-sided (1 α)-level CI for ξ
when applying SRS is
J
n
= [ξ
n
±t
α
S
b,m
/
b].
The asymptotic validity of J
n
can be established by
exploiting the Bahadur representation in (2) for fixed
p
n
= p. An advantage of sectioning over batching
arises from the fact that quantile estimators are gen-
erally biased. While the bias decreases (nonmono-
tonically) to zero as the sample size n increases, it
can be significant for small sample sizes. The bias
of the batching point estimator
¯
ξ
b,m
is determined by
the batch size m = n/b < n, so
¯
ξ
b,m
can be consid-
erably more biased than the overall quantile estima-
tor ξ
n
, which has bias governed by the overall sample
size n. Since the sectioning CI is centered at a less-
biased point ξ
n
, whereas the batching CI is centered
at
¯
ξ
b,m
, the sectioning CI typically has better cover-
age than the batching CI for a fixed overall sample
size n = bm; see the numerical results in Section 4.
3 CONDITIONAL MONTE
CARLO
Now suppose that (X,Y) is a random vector with
joint distribution H. As before, we want to estimate
ξ = F
1
(p), where F again denotes the (marginal)
distribution of X. Let (X
i
,Y
i
), i = 1, 2,... ,n, be a sam-
ple of n i.i.d. pairs from H.
Since F(x) = E[E[I(X x)|Y ]] = E[P(X
x|Y)], a conditional Monte Carlo estimator of F(x)
is
ˆ
F
n
(x) =
1
n
n
i=1
E[I(X
i
x)|Y
i
] =
1
n
n
i=1
G(Y
i
,x), (7)
where G(Y, x) = P(X x|Y). The CMC p-quantile
estimator is
ˆ
ξ
n
=
ˆ
F
1
n
(p). Applying CMC relies crit-
ically on being able to compute G and invert
ˆ
F
n
.
Computing
ˆ
F
1
n
(p) for CMC appears to be more
involved than for SRS or the other variance-reduction
techniques examined in (Chu and Nakayama, 2012).
For example, consider the simple case when (X,Y)
has a bivariate normal distribution with zero marginal
means, unit marginal variances, and correlation ρ.
The conditional distribution of X given Y = y is
N(ρy,1ρ
2
) (e.g., see pp. 167–168 of (Mood et al.,
1974)), so
G(Y,x) = P(X x|Y) = Φ
xρY
p
1ρ
2
!
, (8)
and
ˆ
F
n
(x) =
1
n
n
i=1
Φ
xρY
i
p
1ρ
2
!
.
Identifying
ˆ
ξ
n
such that
ˆ
F
n
(
ˆ
ξ
n
) = p, i.e.,
ˆ
ξ
n
=
ˆ
F
1
n
(p),
can be accomplished using a root-finding algorithm,
e.g., Newton’s method, the secant method or the false-
position method; e.g., see Sections 7.1 and 7.2 of
(Ortega and Rheinboldt, 1987). In contrast to the
secant and false-position methods, Newton’s method
requires computing the derivative of
ˆ
F
n
. Given
Y
1
,Y
2
,... ,Y
n
, note that
ˆ
F
n
(x) is strictly increasing and
differentiable in x, with sample-path derivative
d
dx
ˆ
F
n
(x) =
1
n
n
i=1
(1ρ
2
)
1/2
φ
xρY
i
p
1ρ
2
!
,
where φ is the density of a standard normal.
The following shows that the CMC p
n
-quantile es-
timator
ˆ
F
1
n
(p
n
) satisfies a Bahadur representation.
Theorem 1. If f(ξ) > 0, then for any p
n
= p +
O(n
1/2
), the CMC p
n
-quantile estimator
ˆ
F
1
n
(p
n
)
satisfies
ˆ
F
1
n
(p
n
) = ξ
p
n
ˆ
F
n
(ξ) p
f(ξ)
+ R
n
, with
nR
n
0
(9)
SIMULTECH2014-4thInternationalConferenceonSimulationandModelingMethodologies,Technologiesand
Applications
282
as n , where ξ
p
n
is given in (4).
In particular, the results of Theorem 1 hold for
p
n
= p fixed, in which case ξ
p
n
= ξ. It then fol-
lows from (9) that the CMC p-quantile estimator
ˆ
ξ
n
=
ˆ
F
1
n
(p) satisfies the following CLT:
n(
ˆ
ξ
n
ξ) =
n
f(ξ)
(
ˆ
F
n
(ξ) p) +
nR
n
N(0,τ
2
) (10)
as n , where
τ
2
=
Var[G(Y, ξ)]
f
2
(ξ)
. (11)
The numerator in the right side of (11) arises
since
ˆ
F
n
(ξ) is the sample average of the i.i.d.
G(Y
i
,ξ), i = 1,2,. . .,n. By the well-known variance-
decomposition formula (e.g., see Section 2.10 of
(Ross, 2006)), we have that
p(1 p) = Var[I(X ξ)]
= E[Var[I(X ξ)|Y]] + Var[E[I(X ξ)|Y]]
Var[E[I(X ξ)|Y]] = Var[G(Y,ξ)],
(12)
where the inequality follows from the nonnegativity
of (conditional) variance. Thus, comparing (1) with
(10) and (11), we see that the CMC p-quantile estima-
tor
ˆ
ξ
n
has smaller asymptotic variance than the SRS
p-quantile estimator ξ
n
.
The CLT in (10) can be unfolded to construct an
asymptotically valid confidence interval for ξ if we
can consistently estimate τ
2
in (11). Theorem 1 can
be used to prove that the finite difference
ˆ
λ
n
=
ˆ
F
1
n
(p+ h
n
)
ˆ
F
1
n
(ph
n
)
2h
n
(13)
satisfies
ˆ
λ
n
λ = 1/ f(ξ) as n when the band-
width h
n
= c/
n for any constant c > 0. We can con-
sistently estimate the numerator ψ
2
Var[G(Y, ξ)] in
(11) via
ˆ
ψ
2
n
=
1
n1
n
i=1
h
G(Y
i
,
ˆ
ξ
n
)
¯
G
n
i
2
,
where
¯
G
n
= (1/n)
n
i=1
G(Y
i
,
ˆ
ξ
n
). If G(y,x) is continu-
ous in x for each y, then
¯
G
n
= p since
ˆ
F
n
(
ˆ
ξ
n
) = p. The
proof of the consistency of
ˆ
ψ
2
n
is complicated by the
fact that G(Y
i
,
ˆ
ξ
n
), i = 1, 2,.. .,n, are not i.i.d. because
they all depend on
ˆ
ξ
n
, which in turn is a function of all
Y
i
, i = 1, 2,.. .,n. But this can be handled employing
arguments developed in (Chu and Nakayama, 2012).
Now we can consistently estimate τ using
ˆ
τ
n
=
ˆ
ψ
n
ˆ
λ
n
,
so an asymptotically valid two-sided CI for ξ is
ˆ
I
n
= [
ˆ
ξ
n
±z
α
ˆ
τ
n
/
n]. (14)
As with SRS, choosing an appropriate value for
the bandwidth in the finite-difference
ˆ
λ
n
can be dif-
ficult when applying CMC, and we may instead ap-
ply batching or sectioning to construct a CI for ξ
with CMC. For batching, we divide the G(Y
i
,·), i =
1,2, . ..,n, into b 2 nonoverlapping batches, each
of size m = n/b. As with SRS, a reasonable choice
for the number of batches is b = 10. For each j =
1,2, . ..,b, the jth batch consists of G(Y
( j1)m+i
,·),
i = 1,2,... , m, which we use to compute a CDF es-
timator
ˆ
F
j,m
, with
ˆ
F
j,m
(x) =
1
m
m
i=1
G(Y
( j1)m+i
,x),
and p-quantile estimator
ˆ
ξ
j,m
=
ˆ
F
1
j,m
(p). The b
batch quantile estimators
ˆ
ξ
j,m
, j = 1,2, ..., b, are
i.i.d., and we compute their sample average
˜
ξ
b,m
=
(1/b)
b
j=1
ˆ
ξ
j,m
and sample variance
˜
S
2
b,m
= (1/(b
1))
b
j=1
(
ˆ
ξ
j,m
˜
ξ
b,m
)
2
. The batching CI for ξ when
applying CMC is then
˜
J
b,m
= [
˜
ξ
b,m
±t
α
˜
S
b,m
/
b].
Because of the bias of quantile estimators, it is
often better to apply sectioning instead of batching
when n is small, where we again replace the batching
point estimator
˜
ξ
b,m
with the overall quantile estima-
tor
ˆ
ξ
n
. Define
ˆ
S
2
b,m
= (1/(b 1))
b
j=1
(
ˆ
ξ
j,m
ˆ
ξ
n
)
2
,
and the sectioning CI for ξ when applying CMC is
then
ˆ
J
b,m
= [
ˆ
ξ
n
±t
α
ˆ
S
b,m
/
b].
As with SRS, when applying CMC, the sectioning CI
ˆ
J
b,m
should have better coverage than the batching CI
˜
J
b,m
for fixed overall sample size n = bm.
The following result establishes the asymptotic
validity of the CMC CIs.
Theorem 2. Suppose f(ξ) > 0. Then the following
hold:
(i) P(ξ
˜
J
b,m
) 1 α and P(ξ
ˆ
J
b,m
) 1 α as
m with b 2 fixed.
(ii) If the bandwidth h
n
= cn
1/2
in (13) for any con-
stant c > 0, then P(ξ
ˆ
I
n
) 1α as n .
4 NUMERICAL RESULTS
We next present numerical results from simulation
experiments on the bivariate normal discussed in
Section 3. Recall (X,Y) is bivariate normal, with
marginal means 0, unit marginal variances, and cor-
relation ρ = 0.5. Our goal is to estimate and construct
QuantileEstimationWhenApplyingConditionalMonteCarlo
283
CIs for the p-quantile ξ of X for different values of
p and different sample sizes n. Tables 1 and 2 con-
tain the results when applying simple random sam-
pling and conditional Monte Carlo, respectively, giv-
ing the estimated coverage of nominal 90% CIs for ξ
and the average half widths (AHWs) of the CIs from
10
4
independent replications, where we use different
methods to construct the CIs.
Table 1: Coverage (and average half width) of nominal 90%
confidence intervals for the p-quantile of a standard nor-
mal X when applying simple random sampling (SRS) with
bandwidth h
n
= 0.2n
1/2
and b = 10 batches.
p = 0.8
n Exact FD Batch Section
100 0.900 0.808 0.620 0.903
(0.235) (0.230) (0.235) (0.260)
400 0.897 0.851 0.821 0.909
(0.118) (0.117) (0.125) (0.129)
1600 0.898 0.875 0.876 0.901
(0.059) (0.059) (0.063) (0.064)
6400 0.902 0.877 0.898 0.905
(0.029) (0.028) (0.032) (0.032)
p = 0.95
n Exact FD Batch Section
100 0.902 0.799 0.825 0.861
(0.348) (0.330) (0.330) (0.340)
400 0.902 0.848 0.646 0.900
(0.174) (0.172) (0.171) (0.188)
1600 0.900 0.878 0.830 0.900
(0.087) (0.087) (0.092) (0.095)
6400 0.901 0.891 0.883 0.902
(0.043) (0.044) (0.047) (0.047)
p = 0.99
n Exact FD Batch Section
100 0.926 0.497 0.024 0.762
(0.614) (0.325) (0.330) (0.502)
400 0.905 0.913 0.696 0.841
(0.307) (0.405) (0.267) (0.284)
1600 0.907 0.891 0.907 0.907
(0.154) (0.160) (0.164) (0.168)
6400 0.902 0.893 0.887 0.902
(0.077) (0.077) (0.082) (0.083)
In each table, the column labeled “Exact” con-
tains the results for the CIs in (6) and (14) but where
we replace the finite difference estimator of λ with
its exact value. This method is typically not imple-
mentable in practice since λ is usually unknown, but
we include results for it as a benchmark to which we
compare the others. For the finite difference (FD),
we use the bandwidth h
n
= 0.2/
n in (5) and (13).
When p 1 and n is small, we can have p + h
n
> 1,
Table 2: Coverage (and average half width) of nominal
90% confidence intervals for the p-quantile of X of a bi-
variate normal (X,Y) with ρ = 0.5 when applying condi-
tional Monte Carlo (CMC) with bandwidth h
n
= 0.2n
1/2
and b = 10 batches.
p = 0.8
n Exact FD Batch Section
100 0.896 0.895 0.892 0.898
(0.087) (0.087) (0.093) (0.093)
400 0.899 0.899 0.899 0.898
(0.043) (0.043) (0.047) (0.047)
1600 0.900 0.899 0.897 0.897
(0.021) (0.021) (0.023) (0.023)
6400 0.896 0.896 0.896 0.896
(0.011) (0.011) (0.012) (0.012)
p = 0.95
n Exact FD Batch Section
100 0.889 0.898 0.873 0.895
(0.105) (0.109) (0.102) (0.103)
400 0.897 0.898 0.893 0.898
(0.049) (0.050) (0.052) (0.052)
1600 0.892 0.893 0.895 0.897
(0.024) (0.024) (0.026) (0.026)
6400 0.895 0.895 0.896 0.897
(0.012) (0.012) (0.013) (0.013)
p = 0.99
n Exact FD Batch Section
100 0.888 0.944 0.822 0.886
(0.172) (0.259) (0.114) (0.116)
400 0.893 0.967 0.882 0.894
(0.065) (0.097) (0.060) (0.061)
1600 0.889 0.913 0.893 0.896
(0.029) (0.032) (0.031) (0.031)
6400 0.893 0.898 0.897 0.897
(0.014) (0.015) (0.015) (0.015)
so the finite differences (5) and (13) become unde-
fined since the inverse of the estimated CDF is eval-
uated outside of its domain. In these cases, we re-
place p+ h
n
and ph
n
with q
1
1(1 p)/10 and
q
2
2p 1 + (1 p)/10, respectively, where q
2
is
chosen so that q
1
and q
2
are symmetric around p; the
denominator in the finite difference is then q
1
q
2
.
The columns labeled “Batch” and “Section” are for
batching and sectioning, respectively, with b = 10
batches. Numerical results in (Nakayama, 2014a)
with b = 10 and b = 20 reveal that b = 20 often leads
to poorer coverage than b = 10 for small n.
In general, we see that in both tables, the cover-
ages converge to the nominal level as n gets large,
demonstrating the CIs’ asymptotic validity. When
p 1 and n is small, sectioning generally gives better
coverage than batching because the former centers its
SIMULTECH2014-4thInternationalConferenceonSimulationandModelingMethodologies,Technologiesand
Applications
284
CI at a less-biased point. Sectioning also outperforms
FD in terms of coverage. Comparing FD and Exact
for p = 0.99 and small n, we see that the AHW of
FD typically is quite different from AHW for Exact,
which indicates that in these cases, FD does a poor
job estimating λ, resulting in FD’s poor coverage.
Relative to SRS, CMC reduces the AHW about
60% (resp., 70% and 80%) for p = 0.8 (resp., p =
0.95 and p = 0.99). Thus, the variance reduction from
CMC improves as we consider more extreme quan-
tiles. For each of the smaller values of n, the cov-
erage for each SRS CI (except Exact) worsens as p
increases. While CMC coverage also degrades some-
what as p approaches 1, the impact is much less pro-
nounced. Also, for large n, the slightly wider AHW
for batching and sectioning compared to Exact and
FD arises because the former two methods are based
on a Student t limit, whereas the latter two rely on a
normal limit, which has lighter tails.
5 CONCLUSIONS
We developed an estimator of a quantile ξ using con-
ditional Monte Carlo, which is guaranteed to reduce
asymptotic variance compared to simple random sam-
pling. We established that the CMC quantile estima-
tor satisfies a weak Bahadur representation, which im-
plies a CLT holds. We used these results to produce
three asymptotically valid confidence intervals for ξ
as the sample size n . The CIs are based on batch-
ing, sectioning and a finite difference. Our numerical
results seem to indicate that compared to SRS, CMC
not only reduces variance, but it also leads to CIs with
better coverage. For both SRS and CMC, the sec-
tioning CI has better coverage than the batching and
finite-difference intervals for small n, especially when
p 1. Thus, of the three CIs we proposed, we recom-
mend using sectioning.
ACKNOWLEDGEMENTS
This work has been supported in part by the Na-
tional Science Foundation under Grants No. CMMI-
0926949, CMMI-1200065, and DMS-1331010. Any
opinions, findings, and conclusions or recommenda-
tions expressed in this material are those of the author
and do not necessarily reflect the views of the Na-
tional Science Foundation.
REFERENCES
Asmussen, S. and Glynn, P. (2007). Stochastic Simulation:
Algorithms and Analysis. Springer, New York.
Bahadur, R. R. (1966). A note on quantiles in large samples.
Annals of Mathematical Statistics, 37:577–580.
Billingsley, P. (1995). Probability and Measure. John Wiley
& Sons, New York, third edition.
Bloch, D. A. and Gastwirth, J. L. (1968). On a simple esti-
mate of the reciprocal of the density function. Annals
of Mathematical Statistics, 39:1083–1085.
Bofinger, E. (1975). Estimation of a density function using
order statistics. Australian Journal of Statistics, 17:1
7.
Chu, F. and Nakayama, M. K. (2012). Confidence intervals
for quantiles when applying variance-reduction tech-
niques. ACM Transactions On Modeling and Com-
puter Simulation, 36:Article 7 (25 pages plus 12–page
online–only appendix).
Falk, M. (1986). On the estimation of the quantile density
function. Statistics & Probability Letters, 4:69–73.
Fu, M. C., Hong, L. J., and Hu, J.-Q. (2009). Conditional
Monte Carlo estimation of quantile sensitivities. Man-
agement Science, 55:2019–2027.
Ghosh, J. K. (1971). A new proof of the Bahadur repre-
sentation of quantiles and an application. Annals of
Mathematical Statistics, 42:1957–1961.
Jorion, P. (2007). Value at Risk: The New Benchmark for
Managing Financial Risk, 3rd Edition. McGraw-Hill.
Mood, A. M., Graybill, F. A., and Boes, D. C. (1974). In-
troduction to the Theory of Statistics. McGraw-Hill,
New York, 3rd edition.
Nakayama, M. K. (2014a). Confidence intervals us-
ing sectioning for quantiles when applying variance-
reduction techniques. ACM Transactions on Modeling
and Computer Simulation. To appear.
Nakayama, M. K. (2014b). Efficient quantile estimation
using conditional Monte Carlo. In preparation.
Ortega, J. M. and Rheinboldt, W. C. (1987). terative So-
lution of Nonlinear Equations in Several Variables.
SIAM.
Ross, S. M. (2006). Simulation. Academic Press, San
Diego, CA, fourth edition.
Serfling, R. J. (1980). Approximation Theorems of Mathe-
matical Statistics. John Wiley & Sons, New York.
U.S. Nuclear Regulatory Commission (1989). Best-
estimate calculations of emergency core cooling per-
formance. Nuclear Regulatory Commission Regula-
tory Guide 1.157, U.S. Nuclear Regulatory Commis-
sion, Washington, DC.
U.S. Nuclear Regulatory Commission (2011). Applying
statistics. U.S. Nuclear Regulatory Commission Re-
port NUREG-1475, Rev 1, U.S. Nuclear Regulatory
Commission, Washington, DC.
QuantileEstimationWhenApplyingConditionalMonteCarlo
285