1\chapter{Reliability Methods}\label{uq:reliability}
2
3%This chapter explores local and global reliability methods in greater
4%detail than that provided in the Uncertainty Quantification chapter of
5%the User's Manual.
6
7
8\section{Local Reliability Methods}\label{uq:reliability:local}
9
10Local reliability methods include the Mean Value method and the family
11of most probable point (MPP) search methods.  Each of these methods is
12gradient-based, employing local approximations and/or local
13optimization methods.
14
15
16\subsection{Mean Value}\label{uq:reliability:local:mv}
17
18The Mean Value method (MV, also known as MVFOSM in \cite{Hal00}) is
19the simplest, least-expensive reliability method because it estimates
20the response means, response standard deviations, and all CDF/CCDF
21response-probability-reliability levels from a single evaluation of
22response functions and their gradients at the uncertain variable
23means.  This approximation can have acceptable accuracy when the
24response functions are nearly linear and their distributions are
25approximately Gaussian, but can have poor accuracy in other
26situations.
27
28The expressions for approximate response mean $\mu_g$ and approximate
29response variance $\sigma^2_g$ are
30\begin{eqnarray}
31\mu_g      & = & g(\mu_{\bf x})  \label{eq:mv_mean1} \\
32\sigma^2_g & = & \sum_i \sum_j Cov(i,j) \frac{dg}{dx_i}(\mu_{\bf x})
33                 \frac{dg}{dx_j}(\mu_{\bf x}) \label{eq:mv_std_dev}
34\end{eqnarray}
35where ${\bf x}$ are the uncertain values in the
36space of the original uncertain variables (``x-space''), $g({\bf x})$
37is the limit state function (the response function for which
38probability-response level pairs are needed), and the use of a linear
39Taylor series approximation is evident.
40These two moments are then used for mappings from response target to
41approximate reliability level ($\bar{z} \to \beta$) and from
42reliability target to approximate response level
43($\bar{\beta} \to z$) using
44\begin{eqnarray}
45\bar{z} \rightarrow \beta: & ~ &
46\beta_{\rm CDF} = \frac{\mu_g - \bar{z}}{\sigma_g}, ~~~~~
47\beta_{\rm CCDF} = \frac{\bar{z} - \mu_g}{\sigma_g} \label{eq:mv_ria} \\
48\bar{\beta} \rightarrow z: & ~ &
49z = \mu_g - \sigma_g \bar{\beta}_{\rm CDF}, ~~~~~
50z = \mu_g + \sigma_g \bar{\beta}_{\rm CCDF} \label{eq:mv_pma}
51\end{eqnarray}
52respectively, where $\beta_{\rm CDF}$ and $\beta_{\rm CCDF}$ are the
53reliability indices corresponding to the cumulative and complementary
54cumulative distribution functions (CDF and CCDF), respectively.
55
56With the introduction of second-order limit state information, MVSOSM
57calculates a second-order mean as
58\begin{equation}
59\mu_g = g(\mu_{\bf x}) + \frac{1}{2} \sum_i \sum_j Cov(i,j)
60\frac{d^2g}{dx_i dx_j}(\mu_{\bf x}) \label{eq:mv_mean2}
61\end{equation}
62This is commonly combined with a first-order variance
63(Equation~\ref{eq:mv_std_dev}), since second-order variance involves
64higher order distribution moments (skewness, kurtosis)~\cite{Hal00}
65which are often unavailable.
66
67The first-order CDF probability $p(g \le z)$, first-order
68CCDF probability $p(g > z)$, $\beta_{\rm CDF}$, and $\beta_{\rm CCDF}$ are
69related to one another through
70\begin{eqnarray}
71p(g \le z)  & = & \Phi(-\beta_{\rm CDF})     \label{eq:p_cdf} \\
72p(g > z)    & = & \Phi(-\beta_{\rm CCDF})    \label{eq:p_ccdf} \\
73\beta_{\rm CDF}  & = & -\Phi^{-1}(p(g \le z)) \label{eq:beta_cdf} \\
74\beta_{\rm CCDF} & = & -\Phi^{-1}(p(g > z))   \label{eq:beta_ccdf} \\
75\beta_{\rm CDF}  & = & -\beta_{\rm CCDF}       \label{eq:beta_cdf_ccdf} \\
76p(g \le z)  & = & 1 - p(g > z)             \label{eq:p_cdf_ccdf}
77\end{eqnarray}
78where $\Phi()$ is the standard normal cumulative distribution
79function, indicating the introduction of a Gaussian assumption on the
80output distributions.  A common convention in the literature is to
81define $g$ in such a way that the CDF probability for a response level
82$z$ of zero (i.e., $p(g \le 0)$) is the response metric of interest.
83Dakota is not restricted to this convention and is designed to support
84CDF or CCDF mappings for general response, probability, and
85reliability level sequences.
86
87With the Mean Value method, it is possible to obtain importance
88factors indicating the relative contribution of the input variables to
89the output variance.  The importance factors can be viewed as an
90extension of linear sensitivity analysis combining deterministic
91gradient information with input uncertainty information, \emph{i.e.}
92input variable standard deviations. The accuracy of the importance
93factors is contingent of the validity of the linear Taylor series
94approximation used to approximate the response quantities of interest.
95The importance factors are determined as follows for each of $n$ random
96variables:
97%, where we require uncorrelated input variables:
98\begin{equation}
99  {\rm ImportFactor}_i = \left[ \frac{\sigma_{x_i}}{\sigma_g}
100  \frac{dg}{dx_i}(\mu_{\bf x}) \right]^2, ~~~~ i = 1, \dots, n
101\end{equation}
102where it is evident that these importance factors correspond to the
103diagonal terms in Eq.~\ref{eq:mv_std_dev} normalized by the total
104response variance.  %involve an attribution of the total response
105%variance among the set of input variables.
106In the case where the input variables are correlated
107resulting in off-diagonal terms for the input covariance, we can also
108compute a two-way importance factor as
109\begin{equation}
110  {\rm ImportFactor}_{ij} = 2 \frac{\sigma^2_{x_{ij}}}{\sigma^2_g}
111  \frac{dg}{dx_i}(\mu_{\bf x}) \frac{dg}{dx_j}(\mu_{\bf x}),
112  ~~~~ i = 1, \dots, n; ~~~~ j = 1, \dots, i-1
113\end{equation}
114These two-way factors differ from the Sobol' interaction terms that
115are computed in variance-based decomposition (refer to
116Section~\ref{uq:expansion:vbd}) due to the non-orthogonality of the
117Taylor series basis.  Due to this non-orthogonality, two-way
118importance factors may be negative, and due to normalization by the
119total response variance, the set of importance factors will always sum
120to one.
121
122
123\subsection{MPP Search Methods}\label{uq:reliability:local:mpp}
124
125All other local reliability methods solve an equality-constrained nonlinear
126optimization problem to compute a most probable point (MPP) and then
127integrate about this point to compute probabilities.  The MPP search
128is performed in uncorrelated standard normal space (``u-space'') since
129it simplifies the probability integration: the distance of the MPP
130from the origin has the meaning of the number of input standard
131deviations separating the mean response from a particular response
132threshold.  The transformation from correlated non-normal
133distributions (x-space) to uncorrelated standard normal distributions
134(u-space) is denoted as ${\bf u} = T({\bf x})$ with the reverse
135transformation denoted as ${\bf x} = T^{-1}({\bf u})$.  These
136transformations are nonlinear in general, and possible approaches
137include the Rosenblatt~\cite{Ros52}, Nataf~\cite{Der86}, and
138Box-Cox~\cite{Box64} transformations.  The nonlinear transformations
139may also be linearized, and common approaches for this include the
140Rackwitz-Fiessler~\cite{Rac78} two-parameter equivalent normal and the
141Chen-Lind~\cite{Che83} and Wu-Wirsching~\cite{Wu87} three-parameter
142equivalent normals.  Dakota employs the Nataf nonlinear transformation
143which is suitable for the common case when marginal distributions and
144a correlation matrix are provided, but full joint distributions are
145not known\footnote{If joint distributions are known, then the
146Rosenblatt transformation is preferred.}.  This transformation occurs
147in the following two steps.  To transform between the
148original correlated x-space variables and correlated standard normals
149(``z-space''), a CDF matching condition is applied for each of the
150marginal distributions:
151\begin{equation}
152\Phi(z_i) = F(x_i) \label{eq:trans_zx}
153\end{equation}
154where $F()$ is the cumulative distribution function of the original
155probability distribution.  Then, to transform between correlated
156z-space variables and uncorrelated u-space variables, the Cholesky
157factor ${\bf L}$ of a modified correlation matrix is used:
158\begin{equation}
159{\bf z} = {\bf L} {\bf u} \label{eq:trans_zu}
160\end{equation}
161where the original correlation matrix for non-normals in x-space has
162been modified to represent the corresponding ``warped'' correlation in
163z-space~\cite{Der86}.
164
165The forward reliability analysis algorithm of computing CDF/CCDF
166probability/reliability levels for specified response levels is called
167the reliability index approach (RIA), and the inverse reliability
168analysis algorithm of computing response levels for specified CDF/CCDF
169probability/reliability levels is called the performance measure
170approach (PMA)~\cite{Tu99}.  The differences between the RIA and PMA
171formulations appear in the objective function and equality constraint
172formulations used in the MPP searches.  For RIA, the MPP search for
173achieving the specified response level $\bar{z}$ is formulated as
174computing the minimum distance in u-space from the origin to the
175$\bar{z}$ contour of the limit state response function:
176\begin{eqnarray}
177{\rm minimize}     & {\bf u}^T {\bf u} \nonumber \\
178{\rm subject \ to} & G({\bf u}) = \bar{z} \label{eq:ria_opt}
179\end{eqnarray}
180where ${\bf u}$ is a vector centered at the origin in u-space and
181$g({\bf x}) \equiv G({\bf u})$ by definition.  For PMA, the MPP search
182for achieving the specified reliability level $\bar{\beta}$ or
183first-order probability level $\bar{p}$ is formulated as computing the
184minimum/maximum response function value corresponding to a prescribed
185distance from the origin in u-space:
186\begin{eqnarray}
187{\rm minimize}     & \pm G({\bf u}) \nonumber \\
188{\rm subject \ to} & {\bf u}^T {\bf u} = \bar{\beta}^2 \label{eq:pma_opt}
189\end{eqnarray}
190where $\bar{\beta}$ is computed from $\bar{p}$ using
191Eq.~\ref{eq:beta_cdf} or~\ref{eq:beta_ccdf} in the latter case of a
192prescribed first-order probability level.  For a specified generalized
193reliability level $\bar{\beta^*}$ or second-order probability level
194$\bar{p}$, the equality constraint is reformulated in terms of the
195generalized reliability index:
196\begin{eqnarray}
197{\rm minimize}     & \pm G({\bf u}) \nonumber \\
198{\rm subject \ to} & \beta^*({\bf u}) = \bar{\beta^*} \label{eq:pma2_opt}
199\end{eqnarray}
200where $\bar{\beta^*}$ is computed from $\bar{p}$ using
201Eq.~\ref{eq:gen_beta} (or its CCDF complement) in the latter case of a
202prescribed second-order probability level.
203
204In the RIA case, the optimal MPP solution ${\bf u}^*$ defines the
205reliability index from $\beta = \pm \|{\bf u}^*\|_2$, which in turn
206defines the CDF/CCDF probabilities (using
207Equations~\ref{eq:p_cdf}-\ref{eq:p_ccdf} in the case of first-order
208integration).  The sign of $\beta$ is defined by
209\begin{eqnarray}
210G({\bf u}^*) > G({\bf 0}): \beta_{\rm CDF} < 0, \beta_{\rm CCDF} > 0 \\
211G({\bf u}^*) < G({\bf 0}): \beta_{\rm CDF} > 0, \beta_{\rm CCDF} < 0
212\end{eqnarray}
213\noindent where $G({\bf 0})$ is the median limit state response computed
214at the origin in u-space\footnote{It is not necessary to explicitly compute
215the median response since the sign of the inner product
216$\langle {\bf u}^*, \nabla_{\bf u} G \rangle$
217can be used to determine the orientation of the optimal response with
218respect to the median response.} (where $\beta_{\rm CDF}$ = $\beta_{\rm CCDF}$ = 0
219and first-order $p(g \le z)$ = $p(g > z)$ = 0.5).  In the PMA case, the
220sign applied to $G({\bf u})$ (equivalent to minimizing or maximizing
221$G({\bf u})$) is similarly defined by either $\bar{\beta}$ (for a specified
222reliability or first-order probability level) or from a $\bar{\beta}$
223estimate\footnote{computed by inverting the second-order probability
224relationships described in Section~\ref{uq:reliability:local:mpp:int} at
225the current ${\bf u}^*$ iterate.} computed from $\bar{\beta^*}$ (for a
226specified generalized reliability or second-order probability level)
227\begin{eqnarray}
228\bar{\beta}_{\rm CDF} < 0, \bar{\beta}_{\rm CCDF} > 0: {\rm maximize \ } G({\bf u}) \\
229\bar{\beta}_{\rm CDF} > 0, \bar{\beta}_{\rm CCDF} < 0: {\rm minimize \ } G({\bf u})
230\end{eqnarray}
231where the limit state at the MPP ($G({\bf u}^*)$) defines the desired
232response level result.
233
234\subsubsection{Limit state approximations} \label{uq:reliability:local:mpp:approx}
235
236There are a variety of algorithmic variations that are available for
237use within RIA/PMA reliability analyses.  First, one may select among
238several different limit state approximations that can be used to
239reduce computational expense during the MPP searches.  Local,
240multipoint, and global approximations of the limit state are possible.
241\cite{Eld05} investigated local first-order limit state
242approximations, and \cite{Eld06a} investigated local second-order
243and multipoint approximations.  These techniques include:
244
245\begin{enumerate}
246\item a single Taylor series per response/reliability/probability level
247in x-space centered at the uncertain variable means.  The first-order
248approach is commonly known as the Advanced Mean Value (AMV) method:
249\begin{equation}
250g({\bf x}) \cong g(\mu_{\bf x}) + \nabla_x g(\mu_{\bf x})^T
251({\bf x} - \mu_{\bf x}) \label{eq:linear_x_mean}
252\end{equation}
253and the second-order approach has been named AMV$^2$:
254\begin{equation}
255g({\bf x}) \cong g(\mu_{\bf x}) + \nabla_{\bf x} g(\mu_{\bf x})^T
256({\bf x} - \mu_{\bf x}) + \frac{1}{2} ({\bf x} - \mu_{\bf x})^T
257\nabla^2_{\bf x} g(\mu_{\bf x}) ({\bf x} - \mu_{\bf x})
258\label{eq:taylor2_x_mean}
259\end{equation}
260
261\item same as AMV/AMV$^2$, except that the Taylor series is expanded
262in u-space.  The first-order option has been termed the u-space AMV
263method:
264\begin{equation}
265G({\bf u}) \cong G(\mu_{\bf u}) + \nabla_u G(\mu_{\bf u})^T
266({\bf u} - \mu_{\bf u}) \label{eq:linear_u_mean}
267\end{equation}
268where $\mu_{\bf u} = T(\mu_{\bf x})$ and is nonzero in general, and
269the second-order option has been named the u-space AMV$^2$ method:
270\begin{equation}
271G({\bf u}) \cong G(\mu_{\bf u}) + \nabla_{\bf u} G(\mu_{\bf u})^T
272({\bf u} - \mu_{\bf u}) + \frac{1}{2} ({\bf u} - \mu_{\bf u})^T
273\nabla^2_{\bf u} G(\mu_{\bf u}) ({\bf u} - \mu_{\bf u})
274\label{eq:taylor2_u_mean}
275\end{equation}
276
277\item an initial Taylor series approximation in x-space at the uncertain
278variable means, with iterative expansion updates at each MPP estimate
279(${\bf x}^*$) until the MPP converges.  The first-order option is
280commonly known as AMV+:
281\begin{equation}
282g({\bf x}) \cong g({\bf x}^*) + \nabla_x g({\bf x}^*)^T ({\bf x} - {\bf x}^*)
283\label{eq:linear_x_mpp}
284\end{equation}
285and the second-order option has been named AMV$^2$+:
286\begin{equation}
287g({\bf x}) \cong g({\bf x}^*) + \nabla_{\bf x} g({\bf x}^*)^T
288({\bf x} - {\bf x}^*) + \frac{1}{2} ({\bf x} - {\bf x}^*)^T
289\nabla^2_{\bf x} g({\bf x}^*) ({\bf x} - {\bf x}^*) \label{eq:taylor2_x_mpp}
290\end{equation}
291
292\item same as AMV+/AMV$^2$+, except that the expansions are performed in
293u-space.  The first-order option has been termed the u-space AMV+ method.
294\begin{equation}
295G({\bf u}) \cong G({\bf u}^*) + \nabla_u G({\bf u}^*)^T ({\bf u} - {\bf u}^*)
296\label{eq:linear_u_mpp}
297\end{equation}
298and the second-order option has been named the u-space AMV$^2$+ method:
299\begin{equation}
300G({\bf u}) \cong G({\bf u}^*) + \nabla_{\bf u} G({\bf u}^*)^T
301({\bf u} - {\bf u}^*) + \frac{1}{2} ({\bf u} - {\bf u}^*)^T
302\nabla^2_{\bf u} G({\bf u}^*) ({\bf u} - {\bf u}^*) \label{eq:taylor2_u_mpp}
303\end{equation}
304
305\item a multipoint approximation in x-space. This approach involves a
306Taylor series approximation in intermediate variables where the powers
307used for the intermediate variables are selected to match information at
308the current and previous expansion points.  Based on the
309two-point exponential approximation concept (TPEA, \cite{Fad90}), the
310two-point adaptive nonlinearity approximation (TANA-3, \cite{Xu98})
311approximates the limit state as:
312\begin{equation}
313g({\bf x}) \cong g({\bf x}_2) + \sum_{i=1}^n
314\frac{\partial g}{\partial x_i}({\bf x}_2) \frac{x_{i,2}^{1-p_i}}{p_i}
315(x_i^{p_i} - x_{i,2}^{p_i}) + \frac{1}{2} \epsilon({\bf x}) \sum_{i=1}^n
316(x_i^{p_i} - x_{i,2}^{p_i})^2 \label{eq:tana_g}
317\end{equation}
318
319\noindent where $n$ is the number of uncertain variables and:
320\begin{eqnarray}
321p_i & = & 1 + \ln \left[ \frac{\frac{\partial g}{\partial x_i}({\bf x}_1)}
322{\frac{\partial g}{\partial x_i}({\bf x}_2)} \right] \left/
323\ln \left[ \frac{x_{i,1}}{x_{i,2}} \right] \right. \label{eq:tana_pi_x} \\
324\epsilon({\bf x}) & = & \frac{H}{\sum_{i=1}^n (x_i^{p_i} - x_{i,1}^{p_i})^2 +
325\sum_{i=1}^n (x_i^{p_i} - x_{i,2}^{p_i})^2} \label{eq:tana_eps_x} \\
326H & = & 2 \left[ g({\bf x}_1) - g({\bf x}_2) - \sum_{i=1}^n
327\frac{\partial g}{\partial x_i}({\bf x}_2) \frac{x_{i,2}^{1-p_i}}{p_i}
328(x_{i,1}^{p_i} - x_{i,2}^{p_i}) \right] \label{eq:tana_H_x}
329\end{eqnarray}
330\noindent and ${\bf x}_2$ and ${\bf x}_1$ are the current and previous
331MPP estimates in x-space, respectively.  Prior to the availability of
332two MPP estimates, x-space AMV+ is used.
333
334\item a multipoint approximation in u-space. The u-space TANA-3
335approximates the limit state as:
336\begin{equation}
337G({\bf u}) \cong G({\bf u}_2) + \sum_{i=1}^n
338\frac{\partial G}{\partial u_i}({\bf u}_2) \frac{u_{i,2}^{1-p_i}}{p_i}
339(u_i^{p_i} - u_{i,2}^{p_i}) + \frac{1}{2} \epsilon({\bf u}) \sum_{i=1}^n
340(u_i^{p_i} - u_{i,2}^{p_i})^2 \label{eq:tana_G}
341\end{equation}
342
343\noindent where:
344\begin{eqnarray}
345p_i & = & 1 + \ln \left[ \frac{\frac{\partial G}{\partial u_i}({\bf u}_1)}
346{\frac{\partial G}{\partial u_i}({\bf u}_2)} \right] \left/
347\ln \left[ \frac{u_{i,1}}{u_{i,2}} \right] \right. \label{eq:tana_pi_u} \\
348\epsilon({\bf u}) & = & \frac{H}{\sum_{i=1}^n (u_i^{p_i} - u_{i,1}^{p_i})^2 +
349\sum_{i=1}^n (u_i^{p_i} - u_{i,2}^{p_i})^2} \label{eq:tana_eps_u} \\
350H & = & 2 \left[ G({\bf u}_1) - G({\bf u}_2) - \sum_{i=1}^n
351\frac{\partial G}{\partial u_i}({\bf u}_2) \frac{u_{i,2}^{1-p_i}}{p_i}
352(u_{i,1}^{p_i} - u_{i,2}^{p_i}) \right] \label{eq:tana_H_u}
353\end{eqnarray}
354\noindent and ${\bf u}_2$ and ${\bf u}_1$ are the current and previous
355MPP estimates in u-space, respectively.  Prior to the availability of
356two MPP estimates, u-space AMV+ is used.
357
358\item the MPP search on the original response functions without the
359use of any approximations.  Combining this option with first-order and
360second-order integration approaches (see next section) results in the
361traditional first-order and second-order reliability methods (FORM and
362SORM).
363\end{enumerate}
364
365The Hessian matrices in AMV$^2$ and AMV$^2$+ may be available
366analytically, estimated numerically, or approximated through
367quasi-Newton updates.  The selection between x-space or u-space for
368performing approximations depends on where the approximation will be
369more accurate, since this will result in more accurate MPP estimates
370(AMV, AMV$^2$) or faster convergence (AMV+, AMV$^2$+, TANA).  Since
371this relative accuracy depends on the forms of the limit state $g(x)$
372and the transformation $T(x)$ and is therefore application dependent
373in general, Dakota supports both options.  A concern with
374approximation-based iterative search methods (i.e., AMV+, AMV$^2$+ and
375TANA) is the robustness of their convergence to the MPP.  It is
376possible for the MPP iterates to oscillate or even diverge.  However,
377to date, this occurrence has been relatively rare, and Dakota contains
378checks that monitor for this behavior.  Another concern with TANA is
379numerical safeguarding (e.g., the possibility of raising negative
380$x_i$ or $u_i$ values to nonintegral $p_i$ exponents in
381Equations~\ref{eq:tana_g}, \ref{eq:tana_eps_x}-\ref{eq:tana_G},
382and~\ref{eq:tana_eps_u}-\ref{eq:tana_H_u}).  Safeguarding involves
383offseting negative $x_i$ or $u_i$ and, for potential numerical
384difficulties with the logarithm ratios in Equations~\ref{eq:tana_pi_x}
385and~\ref{eq:tana_pi_u}, reverting to either the linear ($p_i = 1$) or
386reciprocal ($p_i = -1$) approximation based on which approximation has
387lower error in $\frac{\partial g}{\partial x_i}({\bf x}_1)$ or
388$\frac{\partial G}{\partial u_i}({\bf u}_1)$.
389
390\subsubsection{Probability integrations} \label{uq:reliability:local:mpp:int}
391
392The second algorithmic variation involves the integration approach for
393computing probabilities at the MPP, which can be selected to be
394first-order (Equations~\ref{eq:p_cdf}-\ref{eq:p_ccdf}) or second-order
395integration.  Second-order integration involves applying a curvature
396correction~\cite{Bre84,Hoh88,Hon99}.  Breitung applies a correction
397based on asymptotic analysis~\cite{Bre84}:
398\begin{equation}
399p = \Phi(-\beta_p) \prod_{i=1}^{n-1} \frac{1}{\sqrt{1 + \beta_p \kappa_i}}
400\label{eq:p_2nd_breit}
401\end{equation}
402where $\kappa_i$ are the principal curvatures of the limit state
403function (the eigenvalues of an orthonormal transformation of
404$\nabla^2_{\bf u} G$, taken positive for a convex limit state) and
405$\beta_p \ge 0$ (a CDF or CCDF probability correction is selected to
406obtain the correct sign for $\beta_p$).  An alternate correction in
407\cite{Hoh88} is consistent in the asymptotic regime ($\beta_p \to \infty$)
408but does not collapse to first-order integration for $\beta_p = 0$:
409\begin{equation}
410p = \Phi(-\beta_p) \prod_{i=1}^{n-1}
411\frac{1}{\sqrt{1 + \psi(-\beta_p) \kappa_i}} \label{eq:p_2nd_hr}
412\end{equation}
413where $\psi() = \frac{\phi()}{\Phi()}$ and $\phi()$ is the standard
414normal density function.  \cite{Hon99} applies further corrections to
415Equation~\ref{eq:p_2nd_hr} based on point concentration methods.  At
416this time, all three approaches are available within the code, but the
417Hohenbichler-Rackwitz correction is used by default (switching the
418correction is a compile-time option in the source code and has not
419been exposed in the input specification).
420
421\subsubsection{Hessian approximations} \label{sec:hessian}
422
423To use a second-order Taylor series or a second-order integration when
424second-order information ($\nabla^2_{\bf x} g$, $\nabla^2_{\bf u} G$,
425and/or $\kappa$) is not directly available, one can estimate the
426missing information using finite differences or approximate it through
427use of quasi-Newton approximations.  These procedures will often be
428needed to make second-order approaches practical for engineering
429applications.
430
431In the finite difference case, numerical Hessians are commonly computed
432using either first-order forward differences of gradients using
433\begin{equation}
434\nabla^2 g ({\bf x}) \cong
435\frac{\nabla g ({\bf x} + h {\bf e}_i) - \nabla g ({\bf x})}{h}
436\end{equation}
437to estimate the $i^{th}$ Hessian column when gradients are analytically
438available, or second-order differences of function values using
439\begin{equation}
440\begin{array}{l}
441\nabla^2 g ({\bf x}) \cong \frac{g({\bf x} + h {\bf e}_i + h {\bf e}_j) -
442g({\bf x} + h {\bf e}_i - h {\bf e}_j) -
443g({\bf x} - h {\bf e}_i + h {\bf e}_j) +
444g({\bf x} - h {\bf e}_i - h {\bf e}_j)}{4h^2}
445\end{array}
446\end{equation}
447to estimate the $ij^{th}$ Hessian term when gradients are not directly
448available.  This approach has the advantage of locally-accurate
449Hessians for each point of interest (which can lead to quadratic
450convergence rates in discrete Newton methods), but has the
451disadvantage that numerically estimating each of the matrix terms can
452be expensive.
453
454Quasi-Newton approximations, on the other hand, do not reevaluate all
455of the second-order information for every point of interest.  Rather,
456they accumulate approximate curvature information over time using
457secant updates.  Since they utilize the existing gradient evaluations,
458they do not require any additional function evaluations for evaluating
459the Hessian terms.  The quasi-Newton approximations of interest
460include the Broyden-Fletcher-Goldfarb-Shanno (BFGS) update
461\begin{equation}
462{\bf B}_{k+1} = {\bf B}_{k} - \frac{{\bf B}_k {\bf s}_k {\bf s}_k^T {\bf B}_k}
463{{\bf s}_k^T {\bf B}_k {\bf s}_k} +
464\frac{{\bf y}_k {\bf y}_k^T}{{\bf y}_k^T {\bf s}_k} \label{eq:bfgs}
465\end{equation}
466which yields a sequence of symmetric positive definite Hessian
467approximations, and the Symmetric Rank 1 (SR1) update
468\begin{equation}
469{\bf B}_{k+1} = {\bf B}_{k} +
470\frac{({\bf y}_k - {\bf B}_k {\bf s}_k)({\bf y}_k - {\bf B}_k {\bf s}_k)^T}
471{({\bf y}_k - {\bf B}_k {\bf s}_k)^T {\bf s}_k} \label{eq:sr1}
472\end{equation}
473which yields a sequence of symmetric, potentially indefinite, Hessian
474approximations.  ${\bf B}_k$ is the $k^{th}$ approximation to
475the Hessian $\nabla^2 g$, ${\bf s}_k = {\bf x}_{k+1} - {\bf x}_k$ is
476the step and ${\bf y}_k = \nabla g_{k+1} - \nabla g_k$ is the
477corresponding yield in the gradients.  The selection of BFGS versus SR1
478involves the importance of retaining positive definiteness in the
479Hessian approximations; if the procedure does not require it, then
480the SR1 update can be more accurate if the true Hessian is not positive
481definite.  Initial scalings for ${\bf B}_0$ and numerical safeguarding
482techniques (damped BFGS, update skipping) are described in \cite{Eld06a}.
483
484
485\subsubsection{Optimization algorithms}
486
487The next algorithmic variation involves the optimization algorithm
488selection for solving Eqs.~\ref{eq:ria_opt} and~\ref{eq:pma_opt}.  The
489Hasofer-Lind Rackwitz-Fissler (HL-RF) algorithm~\cite{Hal00} is a
490classical approach that has been broadly applied.  It is a
491Newton-based approach lacking line search/trust region globalization,
492and is generally regarded as computationally efficient but
493occasionally unreliable.  Dakota takes the approach of employing
494robust, general-purpose optimization algorithms with provable
495convergence properties.  In particular, we employ the sequential
496quadratic programming (SQP) and nonlinear interior-point (NIP)
497optimization algorithms from the NPSOL~\cite{Gil86} and
498OPT++~\cite{MeOlHoWi07} libraries, respectively.
499
500
501\subsubsection{Warm Starting of MPP Searches}
502
503The final algorithmic variation for local reliability methods involves
504the use of warm starting approaches for improving computational
505efficiency.  \cite{Eld05} describes the acceleration of MPP
506searches through warm starting with approximate iteration increment,
507with $z/p/\beta$ level increment, and with design variable increment.
508Warm started data includes the expansion point and associated response
509values and the MPP optimizer initial guess.  Projections are used when
510an increment in $z/p/\beta$ level or design variables occurs.  Warm
511starts were consistently effective in \cite{Eld05}, with greater
512effectiveness for smaller parameter changes, and are used by default
513in Dakota. %for all computational experiments presented in this paper.
514
515
516\section{Global Reliability Methods}\label{uq:reliability:global}
517
518Local reliability methods, while computationally efficient, have
519well-known failure mechanisms.  When confronted with a limit state
520function that is nonsmooth, local gradient-based optimizers may stall
521due to gradient inaccuracy and fail to converge to an MPP.  Moreover,
522if the limit state is multimodal (multiple MPPs), then a
523gradient-based local method can, at best, locate only one local MPP
524solution.  Finally, a linear (Eqs.~\ref{eq:p_cdf}--\ref{eq:p_ccdf}) or
525parabolic (Eqs.~\ref{eq:p_2nd_breit}--\ref{eq:p_2nd_hr}) approximation
526to the limit state at this MPP may fail to adequately capture the
527contour of a highly nonlinear limit state.  %For these reasons,
528%efficient global reliability analysis (EGRA) is investigated
529%in~\cite{bichon_egra_sdm}.
530
531%Global reliability methods include the efficient global reliability
532%analysis (EGRA) method. Analytical methods of reliability analysis solve a
533%local optimization problem to locate the most probable point of failure (MPP),
534%and then quantify the reliability based on its location and an approximation
535%to the shape of the limit state at this point. Typically, gradient-based
536%solvers are used to solve this optimization problem, which may fail to
537%converge for nonsmooth response functions with unreliable gradients or
538%may converge to only one of several solutions for response functions that
539%possess multiple local optima. In addition to these MPP convergence issues,
540%the evaluated probabilites can be adversely affected by limit state
541%approximations that may be inaccurate. Analysts are then forced
542%to revert to sampling methods, which do not rely on MPP convergence or
543%simplifying approximations to the true shape of the limit state.
544%However, employing such methods is impractical when evaluation of the
545%response function is expensive.
546
547A reliability analysis method that is both efficient when applied to
548expensive response functions and accurate for a response function of
549any arbitrary shape is needed. This section develops such a method
550based on efficient global optimization~\cite{Jon98} (EGO) to the
551search for multiple points on or near the limit state throughout the
552random variable space. By locating multiple points on the limit state,
553more complex limit states can be accurately modeled, resulting in a
554more accurate assessment of the reliability. It should be emphasized
555here that these multiple points exist on a single limit state. Because
556of its roots in efficient global optimization, this method of
557reliability analysis is called efficient global reliability analysis
558(EGRA)~\cite{Bichon2007}.  The following two subsections describe two
559capabilities that are incorporated into the EGRA algorithm: importance
560sampling and EGO.
561
562\subsection{Importance Sampling}\label{uq:reliability:global:ais}
563
564An alternative to MPP search methods is to directly
565perform the probability integration numerically by sampling the
566response function.
567Sampling methods do not rely on a simplifying approximation to the shape
568of the limit state, so they can be more accurate than FORM and SORM, but they
569can also be prohibitively expensive because they generally require a large
570number of response function evaluations.
571Importance sampling methods reduce this expense by focusing the samples in
572the important regions of the uncertain space.
573They do this by centering the sampling density function at the MPP rather
574than at the mean.
575This ensures the samples will lie the region of interest,
576thus increasing the efficiency of the sampling method.
577Adaptive importance sampling (AIS) further improves the efficiency by
578adaptively updating the sampling density function.
579Multimodal adaptive importance sampling~\cite{Dey98,Zou02} is a
580variation of AIS that allows for the use of multiple sampling densities
581making it better suited for cases where multiple sections of the limit state
582are highly probable.
583
584Note that importance sampling methods require that the location of at
585least one MPP be known because it is used to center the initial sampling
586density. However, current gradient-based, local search methods used in
587MPP search may fail to converge or may converge to poor solutions for
588highly nonlinear problems, possibly making these methods inapplicable.
589As the next section describes, EGO is a global optimization method that
590does not depend on the availability of accurate gradient information, making
591convergence more reliable for nonsmooth response functions.
592Moreover, EGO has the ability to locate multiple failure points,
593which would provide multiple starting points and thus a good
594multimodal sampling density for the initial steps of multimodal AIS.
595The resulting Gaussian process model is accurate in the
596vicinity of the limit state, thereby providing an inexpensive surrogate that
597can be used to provide response function samples.
598As will be seen, using EGO to locate multiple points along the limit state,
599and then using the resulting Gaussian process model to provide function
600evaluations in multimodal AIS for the probability integration,
601results in an accurate and efficient reliability analysis tool.
602
603\subsection{Efficient Global Optimization}\label{uq:reliability:global:ego}
604
605Chapter \ref{uq:ego} is now rewritten to support EGO/Bayesian optimization theory.
606
607\subsubsection{Expected Feasibility Function}\label{uq:reliability:global:ego:eff}
608
609The expected improvement function provides an indication of how much the true
610value of the response at a point can be expected to be less than the current
611best solution.
612It therefore makes little sense to apply this to the forward reliability problem
613where the goal is not to minimize the response, but rather to find where it is
614equal to a specified threshold value.
615The expected feasibility function (EFF) is introduced here to provide an
616indication of how well the true value of the response is expected to satisfy
617the equality constraint $G({\bf u})\!=\!\bar{z}$.
618Inspired by the contour estimation work in~\cite{Ran08}, this
619expectation can be calculated in a similar fashion as Eq.~\ref{eq:eif_int}
620by integrating over a region in the immediate vicinity of the threshold value
621$\bar{z}\pm\epsilon$:
622\begin{equation}
623EF\bigl( \hat{G}({\bf u}) \bigr) =
624  \int_{z-\epsilon}^{z+\epsilon}
625    \bigl[ \epsilon - | \bar{z}-G | \bigr] \, \hat{G}({\bf u}) \; dG
626\end{equation}
627\noindent where $G$ denotes a realization of the distribution $\hat{G}$, as
628before.
629Allowing $z^+$ and $z^-$ to denote $\bar{z}\pm\epsilon$, respectively, this
630integral can be expressed analytically as:
631\begin{align}
632EF\bigl( \hat{G}({\bf u}) \bigr) &= \left( \mu_G - \bar{z} \right)
633           \left[ 2 \, \Phi\left( \frac{\bar{z} - \mu_G}{\sigma_G} \right) -
634                       \Phi\left( \frac{  z^-   - \mu_G}{\sigma_G} \right) -
635                       \Phi\left( \frac{  z^+   - \mu_G}{\sigma_G} \right)
636          \right] \notag \\ & \ \ \ \ \ \ \ \ -
637  \sigma_G \left[ 2 \, \phi\left( \frac{\bar{z} - \mu_G}{\sigma_G} \right) \, -
638                       \phi\left( \frac{  z^-   - \mu_G}{\sigma_G} \right) \, -
639                       \phi\left( \frac{  z^+   - \mu_G}{\sigma_G} \right)
640          \right] \notag \\ & \ \ \ \ \ \ \ \ + \ \ \,
641  \epsilon \left[      \Phi\left( \frac{  z^+   - \mu_G}{\sigma_G} \right) -
642                       \Phi\left( \frac{  z^-   - \mu_G}{\sigma_G} \right)
643          \right] \label{eq:eff}
644\end{align}
645where $\epsilon$ is proportional to the standard deviation of the GP
646predictor ($\epsilon\propto\sigma_G$).
647In this case, $z^-$, $z^+$, $\mu_G$, $\sigma_G$, and $\epsilon$ are all functions of the
648location ${\bf u}$, while $\bar{z}$ is a constant.
649Note that the EFF provides the same balance between exploration and
650exploitation as is captured in the EIF.
651Points where the expected value is close to the threshold
652($\mu_G\!\approx\!\bar{z}$) and points with a large uncertainty in the
653prediction will have large expected feasibility values.
654