You are on page 1of 14

Reliability Engineering and System Safety 153 (2016) 96–109

Contents lists available at ScienceDirect

Reliability Engineering and System Safety


journal homepage: www.elsevier.com/locate/ress

A discretization procedure for rare events in Bayesian networks


Kilian Zwirglmaier 1,n, Daniel Straub
Engineering Risk Analysis Group, Technische Universität München, Germany

art ic l e i nf o a b s t r a c t

Article history: Discrete Bayesian networks (BNs) can be effective for risk- and reliability assessments, in which prob-
Received 26 October 2015 ability estimates of (rare) failure events are frequently updated with new information. To solve such
Received in revised form reliability problems accurately in BNs, the discretization of continuous random variables must be per-
7 April 2016
formed carefully. To this end, we develop an efficient discretization scheme, which is based on finding an
Accepted 16 April 2016
Available online 23 April 2016
optimal discretization for the linear approximation of the reliability problem obtained from the First-
Order Reliability Method (FORM). Because the probability estimate should be accurate under all possible
Keywords: future information scenarios, the discretization scheme is optimized with respected to the expected
Bayesian networks posterior error. To simplify application of the method, we establish parametric formulations for efficient
Discretization
discretization of random variables in BNs for reliability problems based on numerical investigations. The
Near-real-time
procedure is implemented into a software prototype. Finally, it is applied to a verification example and an
Structural reliability
Updating application example, the prediction of runway overrun of a landing aircraft.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction among the variables. In the case of discrete BNs, conditional


probability tables (CPTs) quantitatively define the type and
For operational risk and reliability management, it is often strength of the dependence among the variables. The entries of
desirable to compute the probability of a rare event F under the CPT of a variable Y i are the probabilities for each state of Y i
potentially evolving information. Examples include warning sys- conditional on all possible combinations of states of its parents.
tems for natural and technical hazards, or the planning of inspec- For hybrid BNs, which include both discrete and continuous
tion and intervention actions in infrastructure systems. Ideally, this variables, exact inference is available only for two special cases,
is achieved through Bayesian updating of PrðFÞ with the new which are BNs with Gaussian nodes, whose means are linear
information Z to the posterior probability PrðFjZÞ. When physically- functions of their parents, and BNs, whose nodes are defined as a
based or empirical models for predicting the rare event exist, such mixture of truncated basic functions (MoTBFs) [17,18]. Otherwise,
updating is possible with structural reliability methods (SRM) approximate inference algorithms are available for hybrid BNs
[24,26,31]. However, it is often difficult to perform the required based on sampling techniques (e.g. [20,9]). However, these are
computations in near-real-time, due to a lack of efficiency or computationally demanding and not generally suitable for near-
robustness. A modeling and computational framework that does real-time decision support [8]. As an alternative, the continuous
facilitate efficient Bayesian updating is the discrete Bayesian net- random variables can be discretized, which enables the use of
work (BN). Hence it was proposed to combine SRMs with discrete exact inference algorithms that exist for general discrete BNs.
Bayesian networks for near-real-time computations [29,30,7]. These include the variable elimination algorithm [32] and the
BNs are based on directed acyclic graphs (DAGs), to efficiently junction tree algorithm [12,19].
define a joint probability distribution pðY Þ over a random vector Y The size of discrete BNs, and the associated computational
[13,14]. The DAG of a BN, which is often referred to as the quali- effort, increases approximately exponentially with the number of
tative part of a BN, consists of a node for each variable in Y and a discrete states of its nodes, which motivates the development of
set of directed links among nodes representing dependence efficient discretization algorithms. While efficient discretization in
the context of machine learning and BNs in general has been
investigated by multiple researchers [15,5], research on efficient
n
Corresponding author.
discretization in the context of engineering risk analysis or struc-
E-mail addresses: kilian.zwirglmaier@tum.de (K. Zwirglmaier),
straub@tum.de (D. Straub). tural reliability has been limited. In general, it is to be dis-
1
www.era.bgu.tum.de tinguished between static and dynamic discretization. While the

http://dx.doi.org/10.1016/j.ress.2016.04.008
0951-8320/& 2016 Elsevier Ltd. All rights reserved.
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 97

former discretizes the BN a-priori before entering evidence (off- 2. Methodology


line), the latter is based on an iterative scheme that updates the
discretization scheme in function of the evidence (online). 2.1. Structural reliability
Dynamic discretization for risk analysis applications has been
developed mainly by Neil et al. [21], based on the work by Kozlov Since the 1970s structural reliability methods have been devel-
and Koller [16]. The procedure starts with an initial discretization oped and applied in the engineering community to estimate failure
of a hybrid BN, for which an approximate entropy error is calcu- probabilities PrðF Þ of components or systems, based on physical or
lated. If the error complies with a convergence criterion, the cur- empirical models. The performance of engineering components is
rent discretization is accepted. Otherwise the discretization is described by a limit state function (LSF) g ðxÞ, where X ¼ ½X 1 ;…;X n 
iteratively altered, by splitting the intervals with the highest is a vector of basic random variables influencing the performance of
the component. By definition, failure corresponds to g ðxÞ taking
entropy error, until the convergence criterion is fulfilled. The  
non-positive values, i.e. the failure event is F ¼ g ðXÞ r 0 . g ðxÞ
approach is implemented in the software AgenaRisk [1]. Other
includes the physical or engineering model, which is often com-
dynamic discretization algorithms for reliability analysis have
putationally demanding. The probability of failure is calculated by
been proposed, e.g. in [33] for dynamic BNs. The advantage of
integrating the probability density function (PDF) of X, f X ðxÞ, over
dynamic discretization is its flexibility when evidence is entered in
the failure domain:
the BN, i.e. when the model is updated with new observation.
Z
Static discretization has the advantage of being computa-
PrðF Þ ¼ f X ðxÞdx ð1Þ
tionally faster and simple to implement. Some considerations for gðxÞ r 0
static discretization of BNs in reliability applications have been
The formulation can be extended to the reliability of general
presented in [25,29,7]. As pointed out by Friis-Hansen [7], for
systems by defining the failure domain as a combination of series
applications in which extreme events are important, discretiza-
and parallel systems [4]. In the general case, there is no analytical
tion of the distribution tails should be performed with care.
solution to Eq. (1) and the integral is potentially high-dimensional.
Static discretization facilitates a careful representation of these
For this reason, structural reliability methods (SRMs) are applied
tails. However, the accuracy of the static discretization varies to approximate it. These include the first- and the second order
with the available evidence. The difficulty is thus to find a dis- reliability method (FORM and SORM) as well as a large variety of
cretization scheme that is optimal under a wide variety of pos- sampling methods, including importance sampling methods such
terior distributions. as directional importance sampling, and sequential sampling
In this paper we derive a procedure for efficiently performing methods such as subset simulation. These methods are well-
static discretization of continuous reliability problems. An optimal documented in the literature [2,22,3,4].
discretization scheme is sought, which minimizes the expected
approximation error with respect to possible future observations
2.2. First order reliability method (FORM)
(evidence). To solve this optimization problem, we propose to
approximate the reliability problem by the First-Order Reliability To obtain an approximation of the probability of failure through
Method (FORM). Section 2 of the paper describes the proposed FORM, the LSF g ðXÞ is transformed to an equivalent LSF GðUÞ in the
methodology. Section 3 presents numerical parameter studies, and space of uncorrelated standard normal random variables U ¼ ½U 1 ;
simple parametric relations for defining an efficient discretization …;U n  (Fig. 1). The transformation is probability conserving, so that
scheme are derived. In Section 4, the procedure is applied to a set Pr½g ðXÞ r 0 ¼ Pr½GðUÞ r 0 ¼ PrðFÞ. A suitable transformation for
of verification examples and to the computation of the probability this purpose, which is consistent with the BN, is the Rosenblatt
of runway overrun of a landing aircraft. While the theory is transformation [10]. In case all basic random variables are inde-
introduced for problems with only one design point, considera- pendent, this transformation reduces to the marginal transfor-
 1  1
tions regarding problems with multiple design points are given in mations: U i ¼ Φ F X i ðX i Þ , with Φ being the inverse standard
the last verification example and in the discussion. normal CDF.

Fig. 1. Design point and linear approximation of the limit state surface. Left side: original random variable space; right side: standard normal space (from [27].
98 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

The FORM approximation of PrðFÞ is obtained by substituting M i or influencing variables I j . Whenever new evidence on these
the LSF in U-space GðUÞ by a linear function GL ðUÞ, i.e. a first-order variables is available, the BN should be evaluated (in near-real time)
Taylor expansion of GðUÞ. The key idea of FORM is to choose as the utilizing exact BN inference algorithms.
expansion point the so-called design point u , which is the point To enable exact inference algorithms, all continuous random
that minimizes ku k subject to GL ðUÞ r 0. u also known as the variables are discretized. These include the X, and possibly the M i
most likely failure point, as it is the point in the failure domain and I j . In the general case, the computational effort for solving
with the highest probability density. Since all marginal distribu- the BN is a direct function of the CPT size of ‘Component per-
tions of the standard uncorrelated multinormal distribution are formance’. The size of this CPT is 2∏ni¼ 1 ni , where n is the number
standard normal, it can be shown that the FORM probability of of random variables in X, and ni is the number of states used for
failure Pr½GL ðUÞ r 0 is discretizing X i . In this paper we do not describe the discretization
  of random variables M i and I j , since it is typically straightforward
Pr½GL ðUÞ r 0 ¼ Φ  β FORM ð2Þ
and does not contribute significantly to computational perfor-
where Φ is the standard normal CDF and βFORM is the distance mance. The key parameter for computational efficiency and
from the origin to the design point, i.e. βFORM ¼ ku k. The problem accuracy is the discretization scheme for X, which is described in
thus reduces to finding the design point u . If GðUÞ is linear, the Sections 2.5 and 2.6.
FORM solution of the probability of failure is exact, otherwise it is an
approximation, which however is sufficiently accurate in most
2.4. Simplification of BNs through node removal
practical applications with limited numbers of random variables [22].
The linearized LSF GL ðUÞ can be written as
Removing random variables from a BN is one possibility to
GL ðUÞ ¼ βFORM  αT U ð3Þ reduce the computational effort associated with a model. A formal
approach for removing nodes from a BN is described in [30]. In
where α ¼ ½α1 ; …; αn  is the vector of FORM importance mea-
order to decide which nodes to remove from the BN the following
sures. These importance measures are defined as
questions should be considered:
ui
αi ¼ ð4Þ
βFORM  Which random variables are relevant for prediction? (These
where ui
is the ith component of the design point coordinates. include ‘Component performance’.)
The αi ’s take values between  1 and 1, and it is kαk ¼ 1. αi is 0, if  Which random variables can potentially be observed? (These
the uncertainty on U i has no influence on PrðGL ðUÞ r 0Þ, and it is include the measurement variables.)
   Which random variables simplify the modeling of dependen-
1 or  1, if U i is the only random variable affecting Pr g L ðUÞ r 0 .
When the original random variables X i are mutually independent, cies? (These are e.g. common influencing factors such as I 2 in
the αi ’s are readily applicable also in the original space, otherwise Fig. 2.)
the αi ’s can be transformed as described in [3].
 For which random variables is it desirable to explicitly show
their influence on component performance?
2.3. Treatment of a reliability problem in a BN
If a random variable does not belong to any of these categories,
We combine discrete BNs and structural reliability concepts to the corresponding node in the BN can be removed. Since the
facilitate updating of rare event (failure) probabilities under new computational efficiency of the model is governed by the size of
observations. The general problem setting is illustrated in the BN of the CPT of the 'Component state’ node, the primary interest is in
Fig. 2. We here limit the presentation to component reliability pro- removing basic random variables X i from the BN. As a measure for
blems; system problems are considered later. The binary random the relevance of a basic random variable, importance measures αi
variable ‘Component performance’ is described by the LSF gðXÞ. from a FORM analysis may be used. To better understand the
The basic random variables X of the model are included in the relation between αi and X i ’s relevance for prediction, consider a
BN as parents of ‘Component performance’. The nodes M i repre- linearized LSF GL ðUÞ. Following [3], the variance of GL ðUÞ can be
sent measurements of individual random variables X i , and nodes I j decomposed as
represent factors influencing the basic random variables. Depen-    
σ 2GL ¼ ∇G2 α21 þ α22 þ ⋯ þ α2n ð5Þ
dence between the variables in X is modeled either directly by
links among them (here X 2 -X 1 and X 4 -X 5 ) or through common where ∇G denotes the gradient vector of the non-linearized LSF
influencing factors (here I 2 -X 3 and I 2 -X 4 ). The component GðUÞ. From Eq. (5) it is seen that a random variable X i with cor-
performance node can have (multiple) child nodes, which, how- responding αi accounts for α2i U100% of the variance σ 2GL . Therefore,
ever, does not impact the discretization of the reliability problem. observing a random variable X i with αi ¼ 0:1 will reduce the var-
Ultimately, the goal is to predict the component performance, i.e. iance σ 2GL by 1%, whereas observing X j with αj ¼ 0:5 will reduce σ 2GL
PrðFÞ, conditional on other model parameters, such as measurements by 25%.

2.5. Discretization of basic random variables

For ease of presentation, we first consider discretization of


statistically independent basic random variables X, i.e. the special
case of the BN in Fig. 2 in which the X i ’s have no parents.2 The
proposed procedure is extended to the general case of dependent
basic random variables thereafter.

2
In a BN, basic random variables X are independent if they are not connected
through links, if they have no common (unknown) ancestors and if no evidence is
Fig. 2. A general BN including a component reliability problem. available on any of their joint descendants.
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 99

2.5.1. Independent basic random variables the computation of the conditional failure probability following
The situation is illustrated in Fig. 3. The performance of the Eq. (8) is based on the prior probability PrðF jY 1 ¼ 3Þ, the dis-
component depends on n statistically independent random variables cretization introduces an approximation in this case. The error
and is described by a LSF g ðXÞ ¼ gðX 1 ; …; X n Þ. For all basic random occurs only in the intervals that are cut by the limit state surface.
variables X i , corresponding measurements M i can be performed. To In the simple one-dimensional case of Fig. 4, an optimal dis-
obtain an equivalent discrete BN, the continuous X i are replaced by cretization approach would be to discretize the whole outcome
the discrete random variables Y i , and the LSF is replaced by the CPT space in two intervals, one capturing the survival and one the
of component performance conditional on Y ¼ ½Y 1 ;…;Y n . For each failure domain. This discretization would have zero approximation
discrete random variable Y i with ni states 1; 2; …; ni , we define a error. However, already in a two-dimensional case, such a solution
 
discretization scheme Di ¼ d0 ; d1 ; …; dni  1 ; dni consisting of ni þ 1 is not possible. This is illustrated in Fig. 5, where the cells cut by
interval boundaries. The first and the last interval boundaries are the limit state surface are indicated in gray. The failure probability
given by the boundaries of X i 0 s outcome space. conditional on measurements calculated according to Eq. (8) will
Since here the X i , and thus the Y i , have no parents, the PMF of necessarily be an approximation. The approximation error will be
Y i is defined as small, if the contribution of the cells cut by the limit state surface
    (the gray cells in Fig. 5) to the total failure probability is small. An
pY i ðjÞ ¼ F X i dj  F X i dj  1 ; with jϵ½1; …; ni  ð6Þ
efficient discretization will thus limit this contribution with as few
where F X i denotes the cumulative distribution function (CDF) of intervals as possible.
X i . The probability of failure corresponding to the discrete BN in
Fig. 3b can be calculated as 2.5.2. Dependent basic random variables
X
n1 X
nn
      Eqs. (6)–(8) must be adjusted when dependence among the X i ’s
PrðF Þ ¼ … pY 1 y1 …pYn yn U Pr F Y 1 ¼ y1 \ … \ Y n ¼ yn is present, in accordance with the case-specific BN structure.
y1 ¼ 1 yn ¼ 1 However, the principles outlined above for independent X 1 ; …; X n
ð7Þ hold equally for dependent basic random variables: The dis-
Note that the discretization does not introduce any approx- cretization error is a function of the cells cut by the limit state
 function.
imation error here, as long as the conditional Pr F Y 1 ¼ y1 \ … \
Y n ¼ yn Þ is computed exactly. When determining an optimal discretization, we propose in the
Once measurements from the nodes M ¼ ½M 1 ;…;M n  are avail- following to find the FORM approximation of the reliability
able, the conditional failure probability can be calculated as
1 Xn1 Xnn pY 1 ðy1 ÞU pM1 jY 1 ðm1 jy1 Þ … pY n ðyn ÞU pMn jY n ðmn jyn ÞU
PrðFjM ¼ mÞ  … continuous
pM ðmÞ y ¼ 1 y ¼ 1 PrðFjY 1 ¼ y1 \ … \ Y n ¼ yn Þ
1 n

ð8Þ
 
where Pr F Y 1 ¼ y1 ; …; Y n ¼ yn is the conditional probability of
component failure given y1 ; …; yn . If no measurements are avail-
able for some of the basic random variables, the corresponding
likelihood terms p
are simply omitted in Eq. (8).
M j Y j mj yj

While the computation of the unconditional failure probability


following Eq. (7) is exact, the computation of the conditional
failure probability through Eq. (8) is only an approximation. The
reason is that the dependence between the measurement variable
discrete
M i and the ‘Component performance’ variable is not fully captured
in the discrete BN (see also [30]). In Fig. 4, this is illustrated for a
reliability problem with one basic random variable X i . Both the
continuous distribution (Fig. 4a) and the corresponding discretized
distribution (Fig. 4b) are updated correctly after observing M 1 .
However, for Eq.  (8) to be exact, also the conditional failure
probabilities Pr F Y 1 ¼ y1 would need to be updated. This can be
observed in Fig. 4a: in interval Y 1 ¼ 3, which is the one cut by
the limit state surface, the ratio of the probability mass in the
failure domain to that in the safe domain changes from the prior to
the posterior case, i.e. PrðF jY 1 ¼ 3Þ a PrðF jY 1 ¼ 3; M 1 ¼ m1 Þ. Since Fig. 4. Discretization error in 1D.

Fig. 3. Representation of a basic reliability problem with n independent basic random variables in a BN. Left: original problem with continuous basic random variables X i ,
right: discrete BN, in which X i 's are substituted with discrete nodes Y i .
100 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

h
i
~ :
expected preposterior error EM~ e d; M
h
i Z
 
opt
d ¼ arg minEM~ e d; M ~ ~ f M~ ðmÞd
¼ arg min e d; m ~ m ~ ð11Þ
d d ~
M

The optimization is thus based on the computation of an


expected value with respect to the possible measurements
outcomes M ~ before having taken any measurements. This is ana-
logous to a preposterior analysis [23,28]. However, unlike in tra-
ditional preposterior analysis, the objective is not to identify an
optimal action under future available information, but to find the
opt
optimal discretization parameters d . The integral in Eq. (11) is
evaluated through a simple Monte Carlo approach. All M ~ i have the
normal distribution with zero mean and variance 1 þ σ 2ϵ .
The parameters in d describing the discretization scheme are:

Fig. 5. Discretization error in 2D.


1. ni : number of intervals used to discretize each random variable
Ui,
problem, which can readily account for the dependence among the
2. wi : width of the discretization frame in the dimension of U i , and
random variables. Hence, there is no need to distinguish between
3. vi : position of the midpoint of the discretization frame relative
cases with independent or dependent random variables. to the design point

2.6. Efficient discretization These parameters are illustrated in Fig. 6. For a problem with n
basic random variables, the full set of optimization parameters is
2.6.1. Optimal discretization of linear problems in standard normal d ¼ ½w1 ; …; wn ; n1 ; …; nn  1 ; v1 ; …; vn .
space Clearly, the discretization error reduces with increasing ni .
To find an efficient discretization of X, we consider the FORM Because the computational efficiency of the final BN is a direct
solution to the reliability problem. Evaluating the linearized FORM function of the size of the CPT associated with component per-
LSF GL ðUÞ is computationally inexpensive once the design point u formance, which is ∏ni¼ 1 ni , we constrain its size. To this end,
is known. Therefore, it is feasible to find a discretization of U that
  we define cup as the maximum allowed number of parameters of
is optimal for the event GL ðUÞ r 0 through optimization. If GðUÞ the CPT of the component state node. This puts a constraint on the
is not strongly non-linear, this solution will be an efficient dis-
  optimization of Eq. (11):
cretization for GðUÞr0 and,  after a transformation to the ori- n
ginal space, also for g ðXÞ r 0 . ∏ ni r cup ð12Þ
As discussed in Section 2.5.1, the approximation error of the i¼1
discretization is associated with the change from the prior to the The optimization is implemented through a two-level
posterior distribution of the basic random variables. A measure of approach. The optimization of the continuous parameters width
optimality must thus consider possible measurements of X or U. wi and position of the discretization frame vi for all i ¼ 1; …; n is
We consider hypothetical measurements M ~ i (this notation is used
carried out using unconstrained nonlinear optimization for fixed
to distinguish hypothetical measurements from actual measure-
ments M i ¼ mi ) of all standard normal random variables U i with
 
additive measurement error ϵi  N 0; σ ϵi . Since both the prior
distribution and the measurement error are normal distributed,
the likelihood is also normal distributed:
 
M~ i fU i ¼ ui g  N ui ; σ ε ð9Þ
i

The posterior, i.e. the conditional distribution of U i given a


measurement outcome M ~ i¼m~ i , is the normal distribution with
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

1 ~
mean 1 þ σ 2 m i and standard deviation 1  1 þ1σ 2 .
εi εi

We define an error measure based on comparing the true


posterior probability of failure P F jM ðmÞ with the posterior prob-
ability of failure calculated from the discretized U, denoted by
P^ F jM ðd;mÞ . Here, d are the parameters defining the discretization.
The proposed error measure is

log P^ ~ ðd;mÞ ~
~ log 10 P FjM~ ðmÞ
eðd; mÞ ~ ¼ 10 FjM
ð10Þ
~
log 10 P FjM~ ðmÞ

The error measure of Eq. (10) represents a tradeoff between the


absolute and the relative error. It weights the (logarithmic) relative
error by the magnitude of the posterior failure probability. This
ensures that the same relative error is considered worse at a
higher probability level compared to an error at a lower
Fig. 6. Schematic representation of a discretization of a linear 2D reliability pro-
probability level. i i
blem. wi is the distance between interval boundaries d1 and dni  1 . All intervals
A-priori, the measurement outcomes are not known. Hence we between these boundaries are equi-spaced. vi is the position of the midpoint of the
define the optimal discretization as the one that minimizes the discretization frame relative to the design point u in dimension i.
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 101

values of ni . The optimization of the discrete ni is performed the optimal discretization frame coincides with the design point,
through a local search algorithm. Note that the optimization is i.e. vopt
i ¼ 0. The optimal discretization widths wopt i vary sig-
performed offline, i.e. prior to running the BN, hence it does not nificantly with the importance measures αi , as shown in Section
affect the goal of near-real time performance of the BN. Further- 3.1.2. However, the optimal number of intervals nopti is approxi-
more, in Section 3 a heuristic is derived that can replace the time- mately the same for all random variables in all investigated cases,
independent of the αi values, i.e. nopt
1=n
consuming solution of the optimization problem. i ¼ cup .

2.6.2. Efficient discretization of the original random variables X 3.1.1. On why the optimal number of intervals nopt
i is independent of
Since the nodes in the BN represent random variables X in their αi
original outcome space, the discretization schemes, which are To better understand why the number of intervals does not
derived for the corresponding standard normal random variables depend on jαi j (for jαi j that are significantly larger than 0), recall
U, need to be transformed to the X-space. In the case of mutually that an efficient discretization scheme should focus on the area
independent random variables X i , any point on the ith interval around the limit state surface. More precisely, the discretization
boundary in U-space – if transformed – will result in the same error in the posterior case is induced by the cells that are cut by
corresponding ith interval boundary in X-space. This is not the the limit state surface. Exemplarily, Fig. 8 shows a linear problem
case for dependent random variables X i , where a mapping of in standard normal space with two basic random variables U 1 and
the interval boundaries in U-space to X-space will not lead to an U 2 , where α1 ¼ 0:45 and α2 ¼ 0:89. While Fig. 8a shows an optimal
orthogonal discretization scheme in X-space. To preserve ortho- discretization with 5 intervals per dimension, Fig. 8b shows a
gonality throughout the transformation, we propose to represent discretization scheme, where the more important random variable
each interval boundary through a characteristic point and deter- U 2 is discretized with 6 intervals and U 1 with 4 intervals. In both
cases, the discretization frame is centered at the design point. It is
mine the boundary in X-space through a transformation of this
observed that the probability mass of the (gray) cells, associated
point. For transforming the interval boundary of X i , the char-
with the discretization error in the posterior case, is higher in
acteristic point is selected as the design point u , where the ith
Fig. 8b than for the optimal discretion scheme in Fig. 8a. In the
element is substituted by the coordinate of the interval boundary.
example shown here it is 1:4∙10  3 compared to 2:5∙10  4 .
In Fig. 7 this is shown for an example with n ¼ 2 random variables.
For input random variables X i with a value of jαi j close to zero,
the above observations do not hold. Following Section 2.4, these
variables should be removed from the BN prior to discretization.
3. Development of an efficient discretization procedure
3.1.2. Dependence of the optimal discretization width on αi
3.1. Optimization of the FORM approximation In the optimization, it is found that the optimal discretization
width wopt
i varies strongly with the random variable’s importance,
We present the optimal discretization for the FORM approx- expressed through αi . It is reminded that the width wi describes
imation GL ðUÞ for n ¼ 2 and 3 dimensions. Extension to higher the domain in which a fine discretization mesh is applied (Fig. 6).
numbers of random variables is discussed. Because the linear LSF In general, wopt
i decreases with increasing αi . This effect can be
employed in FORM is described only by the reliability index β FORM observed in Fig. 8a, where U 2 is the more important input random
and the vector α of FORM sensitives (Eq. (3)), it facilitates para- variable and it is wopt opt
2 ow1 .
metric studies. A clear relation between wopti and αi can be observed by plot-
Initially, we consider a reliability index β FORM ¼ 4:26, corre- ting the probability mass enclosed by wopt against αi , as shown in
i
sponding to a probability of failure of 10  5 . The standard deviation Fig. 9. The results of Fig. 9 indicate that the probability mass
of the additive measurement error is set to either σ ε ¼ 0:5 or contained within this interval should be a direct function of αi . The
σ ε ¼ 1:0. Different combinations of FORM sensitivity values αi are more important the variable, the finer the discretization around
selected, to investigate their effect on the optimal discretization. In the design point should become. The observed relationship
all investigated cases, we find that the position of the midpoint of between this probability mass and αi follows a clear trend, and a

Fig. 7. Transformation of a discretization scheme from U-space to X-space. To preserve orthogonality each interval boundary in U-space is represented by a characteristic
point. The random variables X 1 and X 2 are Weibull distributed with scale and shape parameter 1 and their correlation is 0.5.
102 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

Fig. 8. Linear problem in standard normal space, with two random variables U 1 and U 2 , where α1 ¼ 0:45 and 0.89. The intervals cut by the limit state surface, i.e. those which
potentially lead to a posterior discretization error are marked in gray.

random variable. The left side of Fig. 10 shows the relation


between the optimal wi and jαi j. The right side of Fig. 10 shows the
same relation, where the wi ’s are scaled as in Fig. 9, i.e. the loga-
rithm of the probability mass enclosed by the outer interval
boundaries is depicted. As in Fig. 9, there is a clear dependence
between the scaled wi values and the jαi j’s. The interval frames
increase with increasing number of random variables.
Secondly, we vary the prior failure probability from 10  3 ðβ ¼
3:1Þ to 10  7 ðβ ¼ 5:2Þ. The results are shown in Fig. 11. Again, a
distinct dependence between the scaled wi values and the jαi j’s is
found. The interval frames decrease with increasing reliability
index (with decreasing failure probability).
Fig. 9. Logarithm of the probability mass enclosed by the discretization frame
plotted against αi . Φ denotes the standard normal CDF and ubi respectively lbi the
last (upper) and the first (lower) interval bound in dimension i.
3.2. Parametric function of optimal discretion frame

function can be fitted (Fig. 9). Neither the dimensionality of the As evident from Figs. 10 and 11, there is a clear dependence of
problem nor the standard deviation of the measurement error the probability mass enclosed by the optimal discretization frame
appear to have an influence on this relation. However, as shown in (with width wi ) on the FORM sensitivity values jαi j. The following
the following section, it is found that the relation does depend on parametric function captures this dependence:
the prior failure probability of the problem (i.e. on β FORM ) and on log ðΦðubi Þ  Φðlbi ÞÞ ¼ a U expðb U jαi jÞ ð13Þ
the number of intervals ni used to discretize the domain.
To facilitate the application in practice and extending the where ubi is the upper and lbi the lower interval boundary in
results to larger numbers of random variables, in Section 3.2 dimension i, such that wi ¼ ubi  lbi . a and b are the parameters of
parametric functions are fitted to the optimization results to the exponential function. This function is depicted in Figs. 10 and 11.
capture the dependency between the optimal discretization width Table 1 shows the parameter values a and b for the different
i and the FORM importance measures αi .
wopt combinations of the prior reliability index β and number of inter-
vals per dimension
3.1.3. Dependence of the optimal discretization on the reliability From the left sides of Figs. 10 and 11, it can be observed that the
index β and the number of discretization cells cup relation between α2i and the optimal wi is fairly diffuse for random
The influence of the prior failure probability and the maximum variables with jαi j o 0:6. Here, the parametric relationship of Eq.
size of the CPT, cup , on the optimal discretization is investigated (13) is less accurate. However, these random variables by defini-
through 10 problems with n ¼ 2 random variables in standard tion have lower importance on the reliability estimate. Hence, the
normal space. The FORM importance measures of the random inaccuracy of Eq. (13) for random variables with jαi j o 0:6 is not
variables are selected between 0:1 and 0:995 and the standard critical, as is confirmed by the numerical investigations performed
deviation of the measurement error is fixed to σ ε ¼ 1:0. We find in the remainder of the paper.
that the optimal discretization frame is generally centered at the The parameter values of Table 1 are derived from two-
design point, i.e. vopt
i ¼ 0, and that the intervals are distributed dimensional problems. In Fig. 9 it is shown that there are no
uniformly among the dimensions. notable differences between two and three dimensions. On this
Firstly, we vary the maximum CPT size cup , i.e. the total number basis, it is hypothesized that the heuristics are applicable also to
of discretization cells. The reliability index is β FORM ¼ 5:2. Fig. 10 problems with higher dimensions. This assumption is furthermore
shows the influence of cup on the resulting width of the dis- supported by the verification examples to be presented in Section 4,
cretization frame wi . Three cases are considered: cup ¼ 25, 100 and where the heuristics are applied also to four-dimensional problems
400. These choices correspond to 5, 10 and 20 intervals for each without any notable deterioration in the results.
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 103

Fig. 10. Optimization results for 10 two-dimensional, linear problems in standard normal space, which are discretized with 5, 10 and 20 intervals per dimension. In all cases
the prior failure probability is 10  7 (β ¼ 5:2). The crosses represent the optimization results. The solid lines are the fitted parametric functions (Eq. (13)). The left-hand side
shows the relation between the width of a discretization frame wi and α2i and the right-hand side shows the relation between the probability mass enclosed by the
discretization frame with width wi and α2i .

3.3. Summary of the proposed procedure c. the width of the discretization frame follows Eq. (13)
6. Transform the discretization scheme to X-space
The steps of the proposed procedure are: 7. Compute the CPTs of the component state node and the basic
random variables using Monte Carlo simulation or Latin
1. Formulate the reliability problem hypercube sampling
2. Set up the corresponding BN
3. Perform a FORM analysis for the reliability problem A MATLAB-based software tool performing these steps is
4. Simplify the BN by removing nodes based on: available for download under www.era.bgu.tum.de/software.
a. their importance for prediction
b. their observability
c. whether or not a node simplifies modeling of dependencies
d. whether or not it is desired to explicitly show a node in the 4. Applications
BN for communication purposes
5. Find the discretization scheme in U-space based on the pro- 4.1. Verification example I
posed heuristics i.e.:
a. the discretization scheme is centered at the design point from For verification purposes, we apply the proposed methodology to
the FORM analysis the discretization of a general limit state with non-normal depen-
b. the same number of intervals is used for each random dent random variables. The approximation error made by this dis-
variable cretization is investigated for different measurement outcomes.
104 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

Fig. 11. Optimization results for 10 two-dimensional, linear problems in standard normal space, which are discretized with 10 intervals per dimension. The prior failure
probabilities are 10  3 (β ¼ 3:1), 10  5 (β ¼ 4:3) and 10  7 (β ¼ 5:2). The crosses represent the optimization results. The solid lines are the fitted parametric functions (Eq. (13)).
The left-hand side shows the relation between the width of a discretization frame wi and α2i and the right-hand side shows the relation between the probability mass
enclosed by the discretization frame with width wi and α2i .

Table 1 The basic random variables are distributed as X 1  LN ð0; 0:5Þ


Parameters a and b of Eq. (13) for β ¼ 3:1, 4.3 and 5.2 as well as 5, 10 and 20 and X 2 ; …; X n  LN ð1; 0:3Þ (values in parenthesis are the para-
intervals per dimension.
meters of the lognormal distribution). The statistical dependence
  among the X i is described through a Gaussian copula model, with
a; ni ¼ 5 ni ¼ 10 ni ¼ 20
b pairwise correlation coefficients ρij . The parameters a and ρij
determine the prior failure probability P F . Measurements M i ¼ mi
  " # " #
β ¼ 3:1  0:28;  1:6 U 10  2 ;  9:8 U 10  4 ; are available for all basic random variables; they are associated
2:9 5:8 8:7 with multiplicative measurement errors ϵi  LNð0; 0:71Þ. In
  " # " #
β ¼ 4:3  0:15;  2:4 U 10  2 ;  2:1 U 10  2 ; Tables 2 and 3, different cases with 3 and 4 random variables are
4:3 6:1 6:2 shown. These cases differ with respect to the prior failure prob-
    " #
β ¼ 5:2  0:36;  0:11;  3:7 U 10  2 ; ability P F , the correlation between the random variables ρij and
3:7 5:0 6:0 the observed measurements m. For each case, a reference solu-
tions P F jM is calculated analytically.
The results in Tables 2 and 3 show that the proposed metho-
Failure is defined through the LSF g ðxÞ: dology for discretization leads to errors in the posterior probability
estimate that are acceptably small for most engineering applica-
n
g ðx Þ ¼ a  ∏ X i ð14Þ tions. (It is reminded that discretization does not lead to a dis-
i¼1 cretization error in the prior case.) As expected, the relative error
is larger when the posterior probability is low, and the absolute
 
i.e., failure corresponds to the event ∏ni¼ 1 X i Z a : error is larger when the posterior probability is high. This follows
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 105

Table 2
Evaluation of the discretization error for different measurement outcomes m, for problems with n ¼ 3 random variables. a is the constant in the LSF, Eq. (14); ρij is the
correlation coefficient between X i and X j for all i a j; P F and P F jM denote the analytically calculated prior and posterior failure probabilities; P^ F jM is the conditional failure
probability calculated with the discrete BN.

a cup ρij PF m P F jM P^ F jM Absolute error Relative error [%]

100 103 0 3.6E  5 ½3:0; 2:9; 2; 9 4.3E  5 4.5E  5 3E  6 6


100 103 0 3.6E  5 ½2:3; 1:1; 2; 1 4.6E  6 5.3E  6 7E  7 14
100 103 0 3.6E  5 ½0:9; 2:4; 0:9 2.8E  7 3.5E  7 7E  8 25
200 153 0:5 1.6E  4 ½1:6; 2:0; 1:2 1.4E  6 1.4E  6 1E  7 4
400 83 0:5 6.4E  6 ½2:6; 3:0; 3:2 8.2E  7 8.9E  7 7E  8 9
400 123 0:5 6.4E  6 ½3:6; 3:3; 4:3 4.9E  6 5.0E  6 1E  9 3

Table 3
Evaluation of the discretization error for different measurement outcomes m. The number of random variables n ¼ 4; a is the constant in the LSF, Eq. (14); ρij is the
correlation coefficient between X i and X j for all i a j; P F and P F jM denote the analytically calculated prior and posterior failure probabilities; P^ FjM is the conditional failure
probability calculated with the discrete BN.

a cup ρij PF m P F jM P^ F jM Absolute error Relative error [%]

400 104 0 1.7E  5 ½2:2; 3:2; 2:4; 3:4 9.5E  6 1.0E  5 9E  7 9


400 104 0 1.7E  5 ½1:6; 1:6; 1:6; 2:0 6.5E  7 7.9E  7 1E  7 21
400 104 0 1.7E  5 ½1:1; 2:3; 1:9; 1:2 2.4E  7 3.0E  7 6E  8 26
600 104 0:5 1.3E  3 ½3:3; 1:7; 2:8; 2:6 4.2E  4 4.3E  4 2E  5 4
800 84 0:5 5.3E  4 ½1:9; 2:0; 1:9; 2:4 1.8E  5 1.9E  5 1E  6 8

from the error measure defined in Eq. (10), which balances the
relative with the absolute error. In addition, the results do not
display any apparent effect of correlation on the accuracy.
To assess the effect of the choice of the number of discretiza-
tion intervals, the failure probability P^ FjM was calculated for a
discretization scheme with up to 20 intervals per RV for the fourth
measurement case in Table 2. The estimated failure probabilities
P^ FjM are plotted together with the exact solution in Fig. 12.

4.2. Verification example II

The failure criterion applied in verification example I (Eq. (14))


leads to a linear LSF in U-space. To verify the accuracy of the Fig. 12. Posterior probability P^ F jM as a function of the number of intervals per
proposed method for problems with non-linear LSFs in U-space, random variable together with the exact (analytical) solution P F jM for the fourth
we additionally investigate the following LSF: measurement case ½1:6; 2:0; 1:2 in Table 2.

X
n
g ðx Þ ¼ a  Xi ð15Þ warning system is developed with the proposed discretization
i¼1 procedure. It provides RWO probabilities conditional on observa-
tions of the landing-weight, the headwind and the approach speed
Again the basic random variables X 1 –X n are distributed as X 1
for different aircraft types and different airports. For a detailed
 LN ð0; 0:5Þ and X 2 ; …; X n  LN ð1; 0:3Þ. Different cases with n ¼ 2,
description of how this problem can be treated in BN framework
3 and 4 random variables are investigated. Measurements M i ¼ mi
we refer to [35].
are available for all basic random variables; associated to these
RWO is the event of the operational landing distance exceeding
measurement are multiplicative measurement errors ϵi  LN
the available runway length (Fig. 13). Correspondingly, a LSF for
ð0; 0:71Þ. For independent random variables X i it is possible to
runway overrun can be defined as:
determine posterior distributions f X i jMi ðxi jmi Þ analytically. The
posterior failure probabilities P FjM , which are used as reference g ðXÞ ¼ Runway n length  Operational n landing
solutions, are calculated through importance sampling with 107 n distanceðXÞ ð16Þ
samples. The results are presented in Table 4.
The results in Table 4 do not differ substantially from with X representing the basic random variables of the problem.
Tables 2 and 3. This indicates that the (weak) non-linearity of the Drees and Holzapfel [6] proposed a model for the operational
LSF function describing failure does not affect the accuracy landing distance required by a landing aircraft, which is applied
significantly. here. The model, as well as the basic random variables X, are
presented in [34], which also includes a detailed description of the
4.3. Runway overrun reliability and sensitivity analysis.
We consider two different airports (AP I and AP II) and two dif-
Runway overrun (RWO) of a landing aircraft is one of the most ferent aircraft types (AC A and AC B). While the aircraft type affects
critical accidents types in civil aviation [11]. A conceptual RWO the landing-weight, the airport affects both the headwind and the
106 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

Table 4
Evaluation of the discretization error for different measurement outcomes m. The problems have n ¼ 2; 3 or 4 random variables; a is the constant in the LSF, Eq. (15); ρij is the
correlation coefficient between X i and X j for all ia j; P F and P F jM denote the prior respectively posterior failure probabilities, which are calculated through importance
sampling with 107 samples; P^ F jM is the conditional failure probability calculated with the discrete BN. Since for correlated basic random variables there is no analytical
solution for the posterior distribution, in these cases the updating of the basic random variables was performed through rejection sampling with 4 5E7 accepted samples.

a cup ρij PF m P FjM P^ F jM Absolute error Rel. error [%]

n¼2
12 102 0 1.3E  5 ½2:8; 4:5 1.4E  5 1.2E  6 2E  6 15
12 102 0 1.3E  5 ½2:3; 2:4 3.3E  6 3.5E  6 2E  7 6
10 122 0 1.7E  4 ½4:0; 3:2 4.0E  4 3.7E  4 3E  5 7
12 102 0:5 1.7E  4 ½2:3; 2:4 4.8E  5 5.0E  5 2E  6 4

n¼3
15 103 0 3.7E  5 ½2:1; 5:6; 5:0 4.8E  5 4.5E  5 4E  6 7
15 103 0 3.7E  5 ½1:1; 3:7; 3:4 1.5E  5 1.8E  5 3E  6 20
13 123 0 5.0E  4 ½3:0; 3:0; 3:0 5.4E  4 5.4E  5 3E  6 1
16 103 0:5 9.1E  4 ½3:0; 6:0; 5:0 1.8E  3 1.9E  3 3E  5 2

n¼4
20 84 0 7.4E  6 ½2:0; 4:0; 3:4; 3:0 4.9E  6 5.6E  6 6E  7 13
17 84 0 3.1E  4 ½1:0; 1:4; 1:2; 2:0 4.5E  5 5.2E  5 7E  6 15
17 124 0 3.1E  4 ½3:1; 2:0; 3:3; 2:4 2.5E  4 2.5E  4 7E  6 3
24 84 0:5 3.7E  4 ½1:0; 1:4; 1:2; 2:0 1.1E  5 1.1E  5 9E  7 8

Table 6
Distribution models for head wind conditional on the airport.

Airport Head wind [kts]

Distribution Mean Std. deviation

I Normal 5:42 5:75


II Normal 6:51 5:75

Table 7
Distribution models for approach speed deviation conditional on the airport.
Fig. 13. Runway definitions.
Airport Approach speed deviation [kts]
Table 5
Distribution Mean Std. deviation
Distribution models for landing weight conditional on the aircraft.

I Gumbel (max) 4:69 4:21


Aircraft Landing weight [t]
II Gumbel (max) 5:63 4:21
Distribution Mean Std. deviation

A Weibull (min) 59:25 1:69


Table 8
B Weibull (min) 64:25 1:69
FORM importance measures αi for each aircraft-airport combination and every
basic random variable in the RWO application.

approach speed. The distribution models for landing-weight, head- Random variable αi Annotation
wind and approach speed deviation at the different airports and
with the different aircraft types are given in Tables 5–7. All other (I/A) (I/B) (II/A) (II/B)
basic random variables of the problem are not affected by the airport
Landing weight [t] 0:09 0:10 0:11 0:09 Modeled
and aircraft type and are as in [34]. Headwind [kts]  0:65  0:61  0:67  0:60 Modeled
Table 8 summarizes the FORM importance measures of all Temperature [°C] 0:03  0:00  0:03  0:03 Not important
random variables X computed for the four combinations of air- Air pressure [hPa] 0:01  0:01  0:01  0:00 Not important
crafts and airports. Touchdown point [m] 0:20 0:16 0:18 0:20 Modeled
Approach speed deviation 0:20 0:21 0:20 0:24 Not observable
[kts]
4.3.1. Selection of relevant random variables Time of spoiler deployment  0:00  0:00 0:01 0:01 Not important
The applied RWO model includes 10 basic random variables. [s]
However, it is sufficient to include only a selection of these Time of breaking initiation 0:70 0:74 0:68 0:73 Not observable
explicitly in the BN. Random variables that are not relevant for the [s]
Time of reverser deploy- 0:03 0:04 0:06 0:05 Not important
prediction of RWO in the considered scenarios can be excluded. ment [s]
This is the case for random variables with a low FORM importance, Time of breaking end [s]  0:02  0:01 0:01 0:02 Not important
whose value does not depend significantly on airport and aircraft
type. Here, all random variables, whose absolute value of the
FORM importance measure jαi j's is smaller than 0:1, are excluded One can additionally exclude random variables that cannot be
(see Table 8). The one exception is landing weight, since its mean measured before the decision on whether to land or not is made.
value is substantially influenced by the aircraft type. This holds for Touchdown point and the time at which the pilot
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 107

Table 9
RWO probabilities for the different airports and aircrafts calculated with the dis-
crete BN pBN , together with solutions calculated by importance sampling around
the design point pDS . The latter have a sampling error with coefficient of variation in
the order of 10%.

AP/AC pBN pDS

I/A 2.0E  7 1.9E  7


I/B 1.0E  6 9.2E  7
II/A 1.3E  7 1.3E  7
II/B 6.9E  7 6.5E  7

Table 10
Probabilities of RWO and corresponding decision on landing, computed with the
BN for different sets of observations.

Case Airport Aircraft Meas. Meas. Meas. Pr ðRWOÞ Landing


LW [t] HW [kts] ASD [kts]

(a) I B 63 0 10:5 2.5E  8 Yes


(b) I A 61  10 5 4.8E  6 No
(c) II B 67 3 0 6.5E  10 Yes
(d) II A 57:5  12 3 1.3E  6 No

Fig. 14. BN structure for a RWO warning system.

solutions, which were calculated by importance sampling


initiates breaking. Since these basic random variables are also not around the design point.
needed to simplify the modeling of dependencies, it is not neces- In Table 10, results obtained with the BN for different hypo-
sary to explicitly model them in the BN, as indicated in Table 8. thetical cases of aircrafts approaching an airport are presented. In
each of these cases, measurements associated with landing
4.3.2. BN model weight, headwind and the approach speed deviation are made. A
The resulting BN of the RWO warning system is shown in threshold on the probability of RWO is used to decide, whether or
Fig. 14. During the aircraft approach, measurements can be not the pilot should continue landing or cancel the landing
obtained for the three basic random variables included in the BN. attempt. Here we assume that up to a RWO probability of 10  6 the
The random variables are discretized separately for each air- pilot should continue landing.
craft–airport combination (joint states of discrete parents) with
8 intervals each, following the proposed discretization procedure.
In a second step, the discretization schemes are merged, i.e. the 5. Discussion
regions of the outcome space, which are discretized with
fine intervals for at least one of the aircraft–airport combinations, When modeling with BNs, it is often necessary or beneficial to
are discretized with the respective fine intervals also in the discretize continuous random variables. When the BN includes
merged discretization scheme. In the end 15 (landing-weight), 10 rare events that are a function of such random variables, the choice
(headwind) and 9 (approach speed deviation) intervals are used to of the discretization scheme is non-trivial. In this contribution, we
discretize the three basic random variables. investigate this discretization based on FORM concepts, and pro-
For all observable quantities, the measurements mi are mod- pose a heuristic procedure for an efficient discretization in these
eled with an additive observation error:
cases. This is based on importance measures αi obtained through a
mi ¼ x i þ ε i ð17Þ FORM analysis, which represent the influence of the uncertainty
associated with a random variable X i .
εi is modeled by a normal distribution with zero mean and The most important finding is that discretization should focus
standard deviation σ εi .
on the area around the most likely failure point (design point),
For the random variable landing weight (at landing time), the
identified by a FORM analysis. Furthermore, we find that optimally
standard deviation of the measurement error is σ ϵLW ¼ 0:34 t. Due
all random variables should be discretized with approximately
to turbulences governing wind speeds, the measurement of
equal numbers of intervals, independent of their importance, as
the head wind speed at the time of the measurement is only an
long as jαi j is not close to zero. The widths of the intervals should
uncertain indicator for the head wind speed at landing time; we
be selected based on the FORM importance αi of the random
model the measurement error with a standard deviation
variables. With increasing importance, the interval width should
σ εHW ¼ 2:88kts. The measurement uncertainty associated with the
approach speed deviation at landing has standard deviation be reduced, leading to finer discretization for larger jαi j. This
σ εASD ¼ 4:21kts. relation is particularly evident for jαi j Z 0:8. We show that it is
49 (Measurement LW), 57 (Measurement HW) and 57 (Measure- possible to fit a parametric function to approximate the relation
ment ASD) intervals are used to discretize the measurement nodes. between jαi j and the optimal width of the region on which the
discretization should focus.
4.3.3. Results This parametric function is used to derive a heuristic procedure
In Table 9, RWO probabilities for the different airports and for finding an efficient discretization. This allows the extrapolation
aircrafts obtained with the discrete BN are compared to of the optimization results to problems with more random
108 K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109

variables. As demonstrated by the verification examples, the  The size of the sub-region of the outcome space of a random
heuristic procedure leads to accurate results. variable X i can be reduced significantly for random variables
This paper is restricted to static discretization. Application of whose corresponding uncertainty is dominating the reliability
the proposed procedure within dynamic discretization (e.g. [21]) problem.
should be investigated. The results of the procedure can serve as  The number of intervals used for discretization should be
an initial discretization scheme, which is iteratively adjusted approximately equal for all basic random variables.
within dynamic discretization. This might strongly enhance the
convergence performance of these algorithms. On this basis, we propose a heuristic that can be used to find an
The gain in computational efficiency resulting from the pro- efficient discretization scheme. In verification examples, this
posed procedure over alternative static discretization approaches heuristic is found to give good accuracy and efficiency.
is problem specific. Some insights can be gained from Fig. 8. A
discretization with intervals of equal width centered at the origin
would require between approximately 2 and 5 times more inter-
vals per random variable to achieve the same accuracy. Because of References
the exponential increase of computational effort with number of
random variables, this leads to a considerable increase in effi- [1] Agena. 2005. Available: 〈http://www.agenarisk.com〉.
ciency. The gain compared to discretization with equal-frequency [2] Au S-K, Beck JL. Estimation of small failure probabilities in high dimensions by
subset simulation. Probab Eng Mech 2001;16:263–77.
intervals is expected to be even higher, since equal frequency [3] Der Kiureghian A. First- and second-order reliability methods. In: Nikolaidis E,
intervals focuses the fine intervals on the region of high prob- Ghiocel DM, Singhal S, editors. Engineering Design Reliability: Handbook. CRC
ability density rather than on the tails of the distribution. Press Inc.; 2005.
[4] Ditlevsen O, Madsen HO. Structural Reliability Methods. John Wiley & Sons;
This paper focuses on component reliability problems, which 2007.
are characterized by a single design point. Nevertheless the [5] Dougherty J, Kohavi R, Mehran S. Supervised and unsupervised discretization
heuristics derived can also be applied to system reliability pro- of continuous features. In: Prieditis A, Russel S, editors. Machine learning:
Proceedings of the twelfth international conference. San Francisco; 1995.
blems. System reliability problems can in general be treated as [6] Drees L, Holzapfel F. Determining and quantifying hazard chains and their
combinations of component reliability problems. Parallel and contribution to incident probabilities in flight operation. In: AIAA modeling
series systems are to be distinguished. For parallel systems dis- and simulation technologies conference. American Institute of Aeronautics
and Astronautics; 2012.
cretization should be performed based on the joint design point of [7] Friis-Hansen A. Bayesian networks as a decision support tool in marine
the problem. For series systems, following the same line of applications. Technical University of Denmark; 2000.
thought as in the runway overrun example, discretization can be [8] Hanea A, Napoles OM, Ababei D. Non-parametric Bayesian networks: improving
theory and reviewing applications. Reliab Eng Syst Saf 2015;144:265–84.
performed separately for each component problem (corresponds [9] Hanea AM, Kurowicka D, Cooke RM. Hybrid method for quantifying and
to the discrete cases i.e. airport–aircraft combinations in the RWO analyzing Bayesian belief nets. Qual Reliab Eng Int 2006;22:709–29.
example). In a second step the discretization schemes can be [10] Hohenbichler M, Rackwitz R. Nonnormal dependent vectors in structural
safety. J Eng Mech 1981;107:1227–38.
merged. In the same way it is possible to apply the heuristic to [11] Iata. Safety report 2012. International Air Transport Association (IATA); 2013.
multi-state components. One can treat each limit state surface [12] Jensen FV, Lauritzen SL, Olesen KG. Bayesian updating in causal probabilistic
(LSF) defining the boundary between two states separately and networks by local computations. Comput Stat Q 1990;4:269–82.
[13] Jensen FV, Nielsen TD. Bayesian networks and decision graphs. New York [u.
merge the discretization schemes afterwards. a.]: Springer; 2007.
The number of basic random variables in a single LSF that can [14] Kjaerulff UB, Madsen AL. Bayesian networks and influence diagrams: a guide
be modeled explicitly in a BN is limited to around 5–8. This is due to construction and analysis. Springer Publishing Company, Incorporated;
2013.
to the exponential growth of the target nodes CPT with increasing [15] Kotsiantis S, Kanellopoulos D. Discretization techniques: a recent survey.
number of parents and is independent of the discretization GESTS Int Trans Comput Sci Eng 2006;32:47–58.
method. Despite this limitation, BNs are applicable to many [16] Kozlov AV, Koller D. Nonuniform dynamic discretization in hybrid networks.
In: Proceedings of the 13th annual conference on Uncertainty in AI (UAI).
practical problems – particularly if one considers that usually not Providence, RI; 1997.
all basic random variables need to be modeled explicitly as nodes, [17] Langseth H, Nielsen TD, Rumi R, Salmerón A. Mixtures of truncated basis
as demonstrated in the presented example. functions. Int J Approx Reason 2012;53:212–27.
[18] Langseth H, Nielsen TD, Rumí R, Salmerón A. Inference in hybrid Bayesian
While in this paper the focus was on the discretization of the networks. Reliab Eng Syst Saf 2009;94:1499–509.
basic random variables, it is straightforward to incorporate the BNs [19] Lauritzen SL, Spiegelhalter DJ. Local computations with probabilities on gra-
discussed into larger models. phical structures and their application to expert systems. J R Stat Soc Ser
B (Methodol) 1988:157–224.
[20] Lerner UN. Hybrid Bayesian networks for reasoning about complex systems.
Stanford University; 2002.
[21] Neil M, Tailor M, Marquez D, Fenton N, Hearty P. Modelling dependable sys-
6. Conclusion
tems using hybrid Bayesian networks. Reliab Eng Syst Saf 2008;93:933–9.
[22] Rackwitz R. Reliability analysis—a review and some perspectives. Struct Saf
We investigate discretization of continuous reliability problems 2001;23:365–95.
[23] Raiffa H, Schlaifer R. Applied statistical decision theory. Boston: Division of
such that they can be treated in a discrete Bayesian network fra-
Research, Graduate School of Business Administration, Harvard University;
mework. Reliability problems with linear LSF in standard normal 1961.
space are considered. These can be seen as FORM approximations [24] Sindel R, Rackwitz R. Problems and Solution Strategies in Reliability Updating.
J Offshore Mech Arct Eng 1998;120:109–14.
of reliability problems. For these linear LSFs, optimal discretization
[25] Straub D. Stochastic modeling of deterioration processes through dynamic
schemes are found, which are optimal with respect to an error Bayesian networks. J Eng Mech 2009;135:1089–99.
measure calculated through a preposterior analysis. Since FORM is [26] Straub D. Reliability updating with equality information. Probab Eng Mech
2011;26:254–8.
known to give good approximations also for most non-linear
[27] Straub D. Engineering risk assessment. Risk – a multidisciplinary introduction.
reliability problem, the resulting discretization schemes are effi- Springer; 2014.
cient also for non-linear LSFs. The main findings presented in this [28] Straub D. Value of information analysis with structural reliability methods.
Struct Saf 2014;49:75–85.
paper are:
[29] Straub D, Der Kiureghian A. Bayesian network enhanced with structural
reliability methods: application. J Eng Mech 2010;136:1259–70.
 An optimal discretization scheme should discretize finely the [30] Straub D, Der Kiureghian A. Bayesian network enhanced with structural
area around the FORM design point. reliability methods: methodology. J Eng Mech 2010;136:1248–58.
K. Zwirglmaier, D. Straub / Reliability Engineering and System Safety 153 (2016) 96–109 109

[31] Straub D, Papaioannou I, Betz W. Bayesian analysis of rare events. J Comput [35] Zwirglmaier K, Straub D. 2015. Discretization of structural reliability pro-
Phys 2016;314:538–56. http://dx.doi.org/10.1016/j.jcp.2016.03.018. blems: an application to runway overrun. In: Proceedings of the 12th Inter-
[32] Zhang NL, Poole D. A simple approach to Bayesian network computations. In: national Conference on Applications of Statistics and Probability in Civil
Proceedings of the tenth Canadian conference on artificial intelligence; 1994. Engineering, ICASP12. Vancouver, Canada.
[33] Zhu J, Collette M. A dynamic discretization method for reliability inference in
dynamic Bayesian networks. Reliab Eng Syst Saf 2015;138:242–52.
[34] Zwirglmaier K, Drees L, Holzapfel F, Straub D. Reliability analysis for runway
overrun using subset simulation. In: Proceedings of ESREL 2014. Wroclaw,
Poland; 2014.

You might also like