1\chapter{Optimization Under Uncertainty (OUU)}\label{ouu}
2
3\section{Reliability-Based Design Optimization (RBDO)}\label{ouu:rbdo}
4
5Reliability-based design optimization (RBDO) methods are used to
6perform design optimization accounting for reliability metrics.  The
7reliability analysis capabilities described in
8Section~\ref{uq:reliability:local} provide a substantial foundation
9for exploring a variety of gradient-based RBDO formulations.
10\cite{Eld05} investigated bi-level, fully-analytic bi-level, and
11first-order sequential RBDO approaches employing underlying
12first-order reliability assessments.  \cite{Eld06a} investigated
13fully-analytic bi-level and second-order sequential RBDO approaches
14employing underlying second-order reliability assessments.  These
15methods are overviewed in the following sections.
16
17\subsection{Bi-level RBDO} \label{ouu:rbdo:bilev}
18
19The simplest and most direct RBDO approach is the bi-level approach in
20which a full reliability analysis is performed for every optimization
21function evaluation.  This involves a nesting of two distinct levels
22of optimization within each other, one at the design level and one at
23the MPP search level.
24
25Since an RBDO problem will typically specify both the $\bar{z}$ level
26and the $\bar{p}/\bar{\beta}$ level, one can use either the RIA or the
27PMA formulation for the UQ portion and then constrain the result in
28the design optimization portion.  In particular, RIA reliability
29analysis maps $\bar{z}$ to $p/\beta$, so RIA RBDO constrains $p/\beta$:
30\begin{eqnarray}
31  {\rm minimize }     & f \nonumber \\
32  {\rm subject \ to } & \beta \ge \bar{\beta} \nonumber \\
33  {\rm or }           & p \le \bar{p} \label{eq:rbdo_ria}
34\end{eqnarray}
35
36\noindent And PMA reliability analysis maps $\bar{p}/\bar{\beta}$ to
37$z$, so PMA RBDO constrains $z$:
38\begin{eqnarray}
39  {\rm minimize }     & f \nonumber \\
40  {\rm subject \ to } & z \ge \bar{z} \label{eq:rbdo_pma}
41\end{eqnarray}
42
43\noindent where $z \ge \bar{z}$ is used as the RBDO constraint for
44a cumulative failure probability (failure defined as $z \le \bar{z}$)
45but $z \le \bar{z}$ would be used as the RBDO constraint for a
46complementary cumulative failure probability (failure defined as $z
47\ge \bar{z}$).  It is worth noting that Dakota is not limited to these
48types of inequality-constrained RBDO formulations; rather, they are
49convenient examples.  Dakota supports general optimization under
50uncertainty mappings~\cite{Eld02} which allow flexible use of
51statistics within multiple objectives, inequality constraints, and
52equality constraints.
53
54An important performance enhancement for bi-level methods is the use
55of sensitivity analysis to analytically compute the design gradients
56of probability, reliability, and response levels.  When design
57variables are separate from the uncertain variables (i.e., they are
58not distribution parameters), then the following first-order
59expressions may be used~\cite{Hoh86,Kar92,All04}:
60\begin{eqnarray}
61\nabla_{\bf d} z           & = & \nabla_{\bf d} g \label{eq:deriv_z} \\
62\nabla_{\bf d} \beta_{cdf} & = & \frac{1}{{\parallel \nabla_{\bf u} G
63\parallel}} \nabla_{\bf d} g \label{eq:deriv_beta} \\
64\nabla_{\bf d} p_{cdf}     & = & -\phi(-\beta_{cdf}) \nabla_{\bf d} \beta_{cdf}
65\label{eq:deriv_p}
66\end{eqnarray}
67where it is evident from Eqs.~\ref{eq:beta_cdf_ccdf}-\ref{eq:p_cdf_ccdf}
68that $\nabla_{\bf d} \beta_{ccdf} = -\nabla_{\bf d} \beta_{cdf}$ and
69$\nabla_{\bf d} p_{ccdf} = -\nabla_{\bf d} p_{cdf}$.  In the case of
70second-order integrations, Eq.~\ref{eq:deriv_p} must be expanded to
71include the curvature correction.  For Breitung's correction
72(Eq.~\ref{eq:p_2nd_breit}),
73\begin{equation}
74\nabla_{\bf d} p_{cdf} = \left[ \Phi(-\beta_p) \sum_{i=1}^{n-1}
75\left( \frac{-\kappa_i}{2 (1 + \beta_p \kappa_i)^{\frac{3}{2}}}
76\prod_{\stackrel{\scriptstyle j=1}{j \ne i}}^{n-1}
77\frac{1}{\sqrt{1 + \beta_p \kappa_j}} \right) -
78\phi(-\beta_p) \prod_{i=1}^{n-1} \frac{1}{\sqrt{1 + \beta_p \kappa_i}}
79\right] \nabla_{\bf d} \beta_{cdf} \label{eq:deriv_p_breit}
80\end{equation}
81where $\nabla_{\bf d} \kappa_i$ has been neglected and $\beta_p \ge 0$
82(see Section~\ref{uq:reliability:local:mpp:int}).  Other approaches assume
83the curvature correction is nearly independent of the design
84variables~\cite{Rac02}, which is equivalent to neglecting the first
85term in Eq.~\ref{eq:deriv_p_breit}.
86
87To capture second-order probability estimates within an RIA RBDO
88formulation using well-behaved $\beta$ constraints, a generalized
89reliability index can be introduced where, similar to Eq.~\ref{eq:beta_cdf},
90\begin{equation}
91\beta^*_{cdf} = -\Phi^{-1}(p_{cdf}) \label{eq:gen_beta}
92\end{equation}
93for second-order $p_{cdf}$.  This reliability index is no longer
94equivalent to the magnitude of ${\bf u}$, but rather is a convenience
95metric for capturing the effect of more accurate probability
96estimates.  The corresponding generalized reliability index
97sensitivity, similar to Eq.~\ref{eq:deriv_p}, is
98\begin{equation}
99\nabla_{\bf d} \beta^*_{cdf} = -\frac{1}{\phi(-\beta^*_{cdf})}
100\nabla_{\bf d} p_{cdf} \label{eq:deriv_gen_beta}
101\end{equation}
102where $\nabla_{\bf d} p_{cdf}$ is defined from Eq.~\ref{eq:deriv_p_breit}.
103Even when $\nabla_{\bf d} g$ is estimated numerically,
104Eqs.~\ref{eq:deriv_z}-\ref{eq:deriv_gen_beta} can be used to avoid
105numerical differencing across full reliability analyses.
106
107When the design variables are distribution parameters of the uncertain
108variables, $\nabla_{\bf d} g$ is expanded with the chain rule and
109Eqs.~\ref{eq:deriv_z} and~\ref{eq:deriv_beta} become
110\begin{eqnarray}
111\nabla_{\bf d} z           & = & \nabla_{\bf d} {\bf x} \nabla_{\bf x} g
112\label{eq:deriv_z_ds} \\
113\nabla_{\bf d} \beta_{cdf} & = & \frac{1}{{\parallel \nabla_{\bf u} G
114\parallel}} \nabla_{\bf d} {\bf x} \nabla_{\bf x} g \label{eq:deriv_beta_ds}
115\end{eqnarray}
116where the design Jacobian of the transformation ($\nabla_{\bf d} {\bf x}$)
117may be obtained analytically for uncorrelated ${\bf x}$ or
118semi-analytically for correlated ${\bf x}$ ($\nabla_{\bf d} {\bf L}$
119is evaluated numerically) by differentiating Eqs.~\ref{eq:trans_zx}
120and~\ref{eq:trans_zu} with respect to the distribution parameters.
121Eqs.~\ref{eq:deriv_p}-\ref{eq:deriv_gen_beta} remain the same as
122before.  For this design variable case, all required information for
123the sensitivities is available from the MPP search.
124
125Since Eqs.~\ref{eq:deriv_z}-\ref{eq:deriv_beta_ds} are derived using
126the Karush-Kuhn-Tucker optimality conditions for a converged MPP, they
127are appropriate for RBDO using AMV+, AMV$^2$+, TANA, FORM, and SORM,
128but not for RBDO using MVFOSM, MVSOSM, AMV, or AMV$^2$.
129
130
131\subsection{Sequential/Surrogate-based RBDO} \label{ouu:rbdo:surr}
132
133An alternative RBDO approach is the sequential approach, in which
134additional efficiency is sought through breaking the nested
135relationship of the MPP and design searches.  The general concept is
136to iterate between optimization and uncertainty quantification,
137updating the optimization goals based on the most recent probabilistic
138assessment results.  This update may be based on safety
139factors~\cite{Wu01} or other approximations~\cite{Du04}.
140
141A particularly effective approach for updating the optimization goals
142is to use the $p/\beta/z$ sensitivity analysis of
143Eqs.~\ref{eq:deriv_z}-\ref{eq:deriv_beta_ds} in combination with local
144surrogate models~\cite{Zou04}.  In \cite{Eld05} and~\cite{Eld06a},
145first-order and second-order Taylor series approximations were
146employed within a trust-region model management framework~\cite{Giu00}
147in order to adaptively manage the extent of the approximations and
148ensure convergence of the RBDO process.  Surrogate models were used
149for both the objective function and the constraints, although the use
150of constraint surrogates alone is sufficient to remove the nesting.
151
152In particular, RIA trust-region surrogate-based RBDO employs surrogate
153models of $f$ and $p/\beta$ within a trust region $\Delta^k$ centered
154at ${\bf d}_c$.  For first-order surrogates:
155\begin{eqnarray}
156  {\rm minimize }     & f({\bf d}_c) + \nabla_d f({\bf d}_c)^T
157({\bf d} - {\bf d}_c) \nonumber \\
158  {\rm subject \ to } & \beta({\bf d}_c) + \nabla_d \beta({\bf d}_c)^T
159({\bf d} - {\bf d}_c) \ge \bar{\beta} \nonumber \\
160  {\rm or }           & p ({\bf d}_c) + \nabla_d p({\bf d}_c)^T
161({\bf d} - {\bf d}_c) \le \bar{p} \nonumber \\
162& {\parallel {\bf d} - {\bf d}_c \parallel}_\infty \le \Delta^k
163\label{eq:rbdo_surr1_ria}
164\end{eqnarray}
165and for second-order surrogates:
166\begin{eqnarray}
167  {\rm minimize }     & f({\bf d}_c) + \nabla_{\bf d} f({\bf d}_c)^T
168({\bf d} - {\bf d}_c)  + \frac{1}{2} ({\bf d} - {\bf d}_c)^T
169\nabla^2_{\bf d} f({\bf d}_c) ({\bf d} - {\bf d}_c) \nonumber \\
170  {\rm subject \ to } & \beta({\bf d}_c) + \nabla_{\bf d} \beta({\bf d}_c)^T
171({\bf d} - {\bf d}_c) + \frac{1}{2} ({\bf d} - {\bf d}_c)^T
172\nabla^2_{\bf d} \beta({\bf d}_c) ({\bf d} - {\bf d}_c) \ge \bar{\beta}
173\nonumber \\
174  {\rm or }           & p ({\bf d}_c) + \nabla_{\bf d} p({\bf d}_c)^T
175({\bf d} - {\bf d}_c) + \frac{1}{2} ({\bf d} - {\bf d}_c)^T
176\nabla^2_{\bf d} p({\bf d}_c) ({\bf d} - {\bf d}_c) \le \bar{p} \nonumber \\
177& {\parallel {\bf d} - {\bf d}_c \parallel}_\infty \le \Delta^k
178\label{eq:rbdo_surr2_ria}
179\end{eqnarray}
180For PMA trust-region surrogate-based RBDO, surrogate models of
181$f$ and $z$ are employed within a trust region $\Delta^k$ centered
182at ${\bf d}_c$.  For first-order surrogates:
183\begin{eqnarray}
184  {\rm minimize }     & f({\bf d}_c) + \nabla_d f({\bf d}_c)^T
185({\bf d} - {\bf d}_c) \nonumber \\
186  {\rm subject \ to } & z({\bf d}_c) + \nabla_d z({\bf d}_c)^T ({\bf d} - {\bf d}_c)
187\ge \bar{z} \nonumber \\
188& {\parallel {\bf d} - {\bf d}_c \parallel}_\infty \le \Delta^k
189\label{eq:rbdo_surr1_pma}
190\end{eqnarray}
191and for second-order surrogates:
192\begin{eqnarray}
193  {\rm minimize }     & f({\bf d}_c) + \nabla_{\bf d} f({\bf d}_c)^T
194({\bf d} - {\bf d}_c) + \frac{1}{2} ({\bf d} - {\bf d}_c)^T
195\nabla^2_{\bf d} f({\bf d}_c) ({\bf d} - {\bf d}_c) \nonumber \\
196  {\rm subject \ to } & z({\bf d}_c) + \nabla_{\bf d} z({\bf d}_c)^T ({\bf d} - {\bf d}_c)
197 + \frac{1}{2} ({\bf d} - {\bf d}_c)^T \nabla^2_{\bf d} z({\bf d}_c)
198({\bf d} - {\bf d}_c) \ge \bar{z} \nonumber \\
199& {\parallel {\bf d} - {\bf d}_c \parallel}_\infty \le \Delta^k
200\label{eq:rbdo_surr2_pma}
201\end{eqnarray}
202where the sense of the $z$ constraint may vary as described
203previously.  The second-order information in
204Eqs.~\ref{eq:rbdo_surr2_ria} and \ref{eq:rbdo_surr2_pma} will
205typically be approximated with quasi-Newton updates.
206
207
208\section{Stochastic Expansion-Based Design Optimization (SEBDO)} \label{ouu:sebdo}
209
210
211\subsection{Stochastic Sensitivity Analysis} \label{ouu:sebdo:ssa}
212
213Section~\ref{uq:expansion:rvsa} describes sensitivity analysis of the
214polynomial chaos expansion with respect to the expansion variables.
215Here we extend this analysis to include sensitivity analysis of
216probabilistic moments with respect to nonprobabilistic (i.e., design
217or epistemic uncertain) variables.
218
219\subsubsection{Local sensitivity analysis: first-order probabilistic expansions} \label{ouu:sebdo:ssa:dvsa_rve}
220
221With the introduction of nonprobabilistic variables $\boldsymbol{s}$
222(for example, design variables or epistemic uncertain variables), a
223polynomial chaos expansion only over the probabilistic variables
224$\boldsymbol{\xi}$ has the functional relationship:
225\begin{equation}
226R(\boldsymbol{\xi}, \boldsymbol{s}) \cong \sum_{j=0}^P \alpha_j(\boldsymbol{s})
227\Psi_j(\boldsymbol{\xi}) \label{eq:R_alpha_s_psi_xi}
228\end{equation}
229
230\noindent For computing sensitivities of response mean and variance,
231the $ij$ indices may be dropped from Eqs.~\ref{eq:mean_pce}
232and~\ref{eq:covar_pce}, simplifying to
233\begin{equation}
234\mu(s) ~=~ \alpha_0(s), ~~~~\sigma^2(s) = \sum_{k=1}^P \alpha^2_k(s) \langle \Psi^2_k \rangle \label{eq:var_pce}
235\end{equation}
236Sensitivities of Eq.~\ref{eq:var_pce} with
237respect to the nonprobabilistic variables are as follows, where
238independence of $\boldsymbol{s}$ and $\boldsymbol{\xi}$ is assumed:
239\begin{eqnarray}
240\frac{d\mu}{ds} &=& \frac{d\alpha_0}{ds} ~~=~~
241%\frac{d}{ds} \langle R \rangle ~~=~~
242\langle \frac{dR}{ds} \rangle \label{eq:dmuR_ds_xi_pce} \\
243\frac{d\sigma^2}{ds} &=& \sum_{k=1}^P \langle \Psi_k^2 \rangle
244\frac{d\alpha_k^2}{ds} ~~=~~
2452 \sum_{k=1}^P \alpha_k \langle \frac{dR}{ds}, \Psi_k \rangle
246\label{eq:dsigR_ds_xi_pce}
247%2 \sigma \frac{d\sigma}{ds} &=& 2
248%\sum_{k=1}^P \alpha_k \frac{d\alpha_k}{ds} \langle \Psi_k^2 \rangle \\
249%\frac{d\sigma}{ds} &=& \frac{1}{\sigma}
250%\sum_{k=1}^P \alpha_k \frac{d}{ds} \langle R, \Psi_k \rangle
251%\label{eq:dsigR_ds_xi_pce}
252\end{eqnarray}
253where
254\begin{equation}
255\frac{d\alpha_k}{ds} = \frac{\langle \frac{dR}{ds}, \Psi_k \rangle}
256{\langle \Psi^2_k \rangle} \label{eq:dalpha_k_ds}
257\end{equation}
258has been used.  Due to independence, the coefficients calculated in
259Eq.~\ref{eq:dalpha_k_ds} may be interpreted as either the derivatives
260of the expectations or the expectations of the derivatives, or more
261precisely, the nonprobabilistic sensitivities of the chaos
262coefficients for the response expansion or the chaos coefficients of
263an expansion for the nonprobabilistic sensitivities of the response.
264The evaluation of integrals involving $\frac{dR}{ds}$ extends the data
265requirements for the PCE approach to include response sensitivities at
266each of the sampled points.% for the quadrature, sparse grid, sampling,
267%or point collocation coefficient estimation approaches.
268The resulting expansions are valid only for a particular set of
269nonprobabilistic variables and must be recalculated each time the
270nonprobabilistic variables are modified.
271
272Similarly for stochastic collocation,
273\begin{equation}
274R(\boldsymbol{\xi}, \boldsymbol{s}) \cong \sum_{k=1}^{N_p} r_k(\boldsymbol{s})
275\boldsymbol{L}_k(\boldsymbol{\xi}) \label{eq:R_r_s_L_xi}
276\end{equation}
277leads to
278\begin{eqnarray}
279\mu(s) &=& \sum_{k=1}^{N_p} r_k(s) w_k, ~~~~\sigma^2(s) ~=~ \sum_{k=1}^{N_p} r^2_k(s) w_k - \mu^2(s) \label{eq:var_sc} \\
280\frac{d\mu}{ds} &=& %\frac{d}{ds} \langle R \rangle ~~=~~
281%\sum_{k=1}^{N_p} \frac{dr_k}{ds} \langle \boldsymbol{L}_k \rangle ~~=~~
282\sum_{k=1}^{N_p} w_k \frac{dr_k}{ds} \label{eq:dmuR_ds_xi_sc} \\
283\frac{d\sigma^2}{ds} &=& \sum_{k=1}^{N_p} 2 w_k r_k \frac{dr_k}{ds}
284- 2 \mu \frac{d\mu}{ds}
285~~=~~ \sum_{k=1}^{N_p} 2 w_k (r_k - \mu) \frac{dr_k}{ds}
286\label{eq:dsigR_ds_xi_sc}
287\end{eqnarray}
288%based on differentiation of Eqs.~\ref{eq:mean_sc}-\ref{eq:covar_sc}.
289
290\subsubsection{Local sensitivity analysis: zeroth-order combined expansions} \label{ouu:sebdo:ssa:dvsa_cve}
291
292Alternatively, a stochastic expansion can be formed over both
293$\boldsymbol{\xi}$ and $\boldsymbol{s}$.  Assuming a bounded
294design domain $\boldsymbol{s}_L \le \boldsymbol{s} \le
295\boldsymbol{s}_U$ (with no implied probability content), a Legendre
296chaos basis would be appropriate for each of the dimensions in
297$\boldsymbol{s}$ within a polynomial chaos expansion.
298\begin{equation}
299R(\boldsymbol{\xi}, \boldsymbol{s}) \cong \sum_{j=0}^P \alpha_j
300\Psi_j(\boldsymbol{\xi}, \boldsymbol{s}) \label{eq:R_alpha_psi_xi_s}
301\end{equation}
302
303\noindent In this case, design sensitivities for the mean and variance
304do not require response sensitivity data, but this comes at the cost
305of forming the PCE over additional dimensions.  For this combined
306variable expansion, the mean and variance are evaluated by performing
307the expectations over only the probabilistic expansion variables,
308which eliminates the polynomial dependence on $\boldsymbol{\xi}$,
309leaving behind the desired polynomial dependence of the moments on
310$\boldsymbol{s}$:
311\begin{eqnarray}
312\mu_R(\boldsymbol{s}) &=& \sum_{j=0}^P \alpha_j \langle \Psi_j(\boldsymbol{\xi},
313\boldsymbol{s}) \rangle_{\boldsymbol{\xi}} \label{eq:muR_comb_pce} \\
314\sigma^2_R(\boldsymbol{s}) &=& \sum_{j=0}^P \sum_{k=0}^P \alpha_j \alpha_k
315\langle \Psi_j(\boldsymbol{\xi}, \boldsymbol{s}) \Psi_k(\boldsymbol{\xi},
316\boldsymbol{s}) \rangle_{\boldsymbol{\xi}} ~-~ \mu^2_R(\boldsymbol{s})
317\label{eq:sigR_comb_pce}
318\end{eqnarray}
319The remaining polynomials may then be differentiated with respect to
320$\boldsymbol{s}$. % as in Eqs.~\ref{eq:dR_dxi_pce}-\ref{eq:deriv_prod_pce}.
321In this approach, the combined PCE is valid for the full design
322variable range ($\boldsymbol{s}_L \le \boldsymbol{s} \le \boldsymbol{s}_U$)
323and does not need to be updated for each change in nonprobabilistic variables,
324although adaptive localization techniques (i.e., trust region model
325management approaches) can be employed when improved local accuracy of
326the sensitivities is required.
327% Q: how is TR ratio formed if exact soln can't be evaluated?
328% A: if objective is accuracy over a design range, then truth is PCE/SC
329%    at a single design point!  -->>  Can use first-order corrections based
330%    on the 2 different SSA approaches!  This is a multifidelity SBO using
331%    HF = probabilistic expansion, LF = Combined expansion. Should get data reuse.
332
333Similarly for stochastic collocation,
334\begin{equation}
335R(\boldsymbol{\xi}, \boldsymbol{s}) \cong \sum_{j=1}^{N_p} r_j
336\boldsymbol{L}_j(\boldsymbol{\xi}, \boldsymbol{s}) \label{eq:R_r_L_xi_s}
337\end{equation}
338leads to
339\begin{eqnarray}
340\mu_R(\boldsymbol{s}) &=& \sum_{j=1}^{N_p} r_j \langle
341\boldsymbol{L}_j(\boldsymbol{\xi}, \boldsymbol{s}) \rangle_{\boldsymbol{\xi}}
342\label{eq:muR_both_sc} \\
343\sigma^2_R(\boldsymbol{s}) &=& \sum_{j=1}^{N_p} \sum_{k=1}^{N_p} r_j r_k
344\langle \boldsymbol{L}_j(\boldsymbol{\xi}, \boldsymbol{s})
345\boldsymbol{L}_k(\boldsymbol{\xi}, \boldsymbol{s}) \rangle_{\boldsymbol{\xi}}
346~-~ \mu^2_R(\boldsymbol{s}) \label{eq:sigR_both_sc}
347\end{eqnarray}
348where the remaining polynomials not eliminated by the expectation over
349$\boldsymbol{\xi}$ are again differentiated with respect to $\boldsymbol{s}$.
350
351\subsubsection{Inputs and outputs} \label{ouu:sebdo:ssa:io}
352
353There are two types of nonprobabilistic variables for which
354sensitivities must be calculated: ``augmented,'' where the
355nonprobabilistic variables are separate from and augment the
356probabilistic variables, and ``inserted,'' where the nonprobabilistic
357variables define distribution parameters for the probabilistic
358variables.  %While one could artificially augment the dimensionality of
359%a combined variable expansion approach with inserted nonprobabilistic
360%variables, this is not currently explored in this work.  Thus, any
361Any inserted nonprobabilistic variable sensitivities must be handled
362using Eqs.~\ref{eq:dmuR_ds_xi_pce}-\ref{eq:dsigR_ds_xi_pce} and
363Eqs.~\ref{eq:dmuR_ds_xi_sc}-\ref{eq:dsigR_ds_xi_sc} where
364$\frac{dR}{ds}$ is calculated as $\frac{dR}{dx} \frac{dx}{ds}$ and
365$\frac{dx}{ds}$ is the Jacobian of the variable transformation
366${\bf x} = T^{-1}(\boldsymbol{\xi})$ with respect to the inserted
367nonprobabilistic variables.  In addition, parameterized polynomials
368(generalized Gauss-Laguerre, Jacobi, and numerically-generated
369polynomials) may introduce a $\frac{d\Psi}{ds}$ or
370$\frac{d\boldsymbol{L}}{ds}$ dependence for inserted $s$ that will
371introduce additional terms in the sensitivity expressions.
372% TO DO: discuss independence of additional nonprobabilistic dimensions:
373% > augmented are OK.
374% > inserted rely on the fact that expansion variables \xi are _standard_
375%   random variables.
376% Special case: parameterized orthogonal polynomials (gen Laguerre, Jacobi)
377% can be differentiated w.r.t. their {alpha,beta} distribution parameters.
378% However, the PCE coefficients are likely also fns of {alpha,beta}.  Therefore,
379% the approach above is correct conceptually but is missing additional terms
380% resulting from the polynomial dependence.
381% NEED TO VERIFY PCE EXPANSION DERIVATIVES FOR PARAMETERIZED POLYNOMIALS!
382
383While moment sensitivities directly enable robust design optimization
384and interval estimation formulations which seek to control or bound
385response variance, control or bounding of reliability requires
386sensitivities of tail statistics.  In this work, the sensitivity of
387simple moment-based approximations to cumulative distribution function
388(CDF) and complementary cumulative distribution function (CCDF)
389mappings (Eqs.~\ref{eq:mv_ria}--\ref{eq:mv_pma}) are employed for this
390purpose, such that it is straightforward to form approximate design
391sensitivities of reliability index $\beta$ (forward reliability
392mapping $\bar{z} \rightarrow \beta$) or response level $z$ (inverse
393reliability mapping $\bar{\beta} \rightarrow z$) from the moment
394design sensitivities and the specified levels $\bar{\beta}$ or
395$\bar{z}$.
396%From here, approximate design sensitivities of probability levels may
397%also be formed given a probability expression (such as $\Phi(-\beta)$)
398%for the reliability index.  The current alternative of numerical
399%design sensitivities of sampled probability levels would employ fewer
400%simplifying approximations, but would also be much more expensive to
401%compute accurately and is avoided for now.  Future capabilities for
402%analytic probability sensitivities could be based on Pearson/Johnson
403%model for analytic response PDFs or
404%sampling sensitivity approaches. % TO DO: cite
405%
406%Extending beyond these simple approaches to support probability and
407%generalized reliability metrics is a subject of current work~\cite{mao2010}.
408
409
410\subsection{Optimization Formulations} \label{ouu:sebdo:form}
411
412Given the capability to compute analytic statistics of the response
413along with design sensitivities of these statistics, Dakota supports
414bi-level, sequential, and multifidelity approaches for optimization
415under uncertainty (OUU). %for reliability-based design and robust design.
416The latter two approaches apply surrogate modeling approaches (data
417fits and multifidelity modeling) to the uncertainty analysis and then
418apply trust region model management to the optimization process.
419
420\subsubsection{Bi-level SEBDO} \label{ouu:sebdo:form:bilev}
421
422The simplest and most direct approach is to employ these analytic
423statistics and their design derivatives from
424Section~\ref{ouu:sebdo:ssa} directly within an optimization loop.
425This approach is known as bi-level OUU, since there is an inner level
426uncertainty analysis nested within an outer level optimization.
427
428Consider the common reliability-based design example of a deterministic
429objective function with a reliability constraint:
430\begin{eqnarray}
431  {\rm minimize }     & f \nonumber \\
432  {\rm subject \ to } & \beta \ge \bar{\beta} \label{eq:rbdo}
433\end{eqnarray}
434where $\beta$ is computed relative to a prescribed threshold response
435value $\bar{z}$ (e.g., a failure threshold) and is constrained by a
436prescribed reliability level $\bar{\beta}$ (minimum allowable
437reliability in the design), and is either a CDF or CCDF index
438depending on the definition of the failure domain (i.e., defined from
439whether the associated failure probability is cumulative, $p(g \le
440\bar{z})$, or complementary cumulative, $p(g > \bar{z})$).
441
442Another common example is robust design in which the
443constraint enforcing a reliability lower-bound has been replaced with
444a constraint enforcing a variance upper-bound $\bar{\sigma}^2$ (maximum
445allowable variance in the design):
446\begin{eqnarray}
447  {\rm minimize }     & f \nonumber \\
448  {\rm subject \ to } & \sigma^2 \le \bar{\sigma}^2 \label{eq:rdo}
449\end{eqnarray}
450
451Solving these problems using a bi-level approach involves computing
452$\beta$ and $\frac{d\beta}{d\boldsymbol{s}}$ for
453Eq.~\ref{eq:rbdo} or $\sigma^2$ and $\frac{d\sigma^2}{d\boldsymbol{s}}$
454for Eq.~\ref{eq:rdo} for each set of design variables $\boldsymbol{s}$
455passed from the optimizer.  This approach is supported for both
456probabilistic and combined expansions using PCE and SC.
457
458\subsubsection{Sequential/Surrogate-Based SEBDO} \label{ouu:sebdo:form:surr}
459
460An alternative OUU approach is the sequential approach, in which
461additional efficiency is sought through breaking the nested
462relationship of the UQ and optimization loops.  The general concept is
463to iterate between optimization and uncertainty quantification,
464updating the optimization goals based on the most recent uncertainty
465assessment results.  This approach is common with the reliability
466methods community, for which the updating strategy may be based on
467safety factors~\cite{Wu01} or other approximations~\cite{Du04}.
468
469A particularly effective approach for updating the optimization goals
470is to use data fit surrogate models, and in particular, local Taylor
471series models allow direct insertion of stochastic sensitivity
472analysis capabilities.  In Ref.~\cite{Eld05}, first-order Taylor
473series approximations were explored, and in Ref.~\cite{Eld06a},
474second-order Taylor series approximations are investigated.  In both
475cases, a trust-region model management framework~\cite{Eld06b} is
476used to adaptively manage the extent of the approximations and ensure
477convergence of the OUU process.  Surrogate models are used for both
478the objective and the constraint functions, although the use of
479surrogates is only required for the functions containing statistical
480results; deterministic functions may remain explicit is desired.
481
482In particular, trust-region surrogate-based optimization for
483reliability-based design employs surrogate models of $f$ and $\beta$
484within a trust region $\Delta^k$ centered at ${\bf s}_c$:
485\begin{eqnarray}
486  {\rm minimize }     & f({\bf s}_c) + \nabla_s f({\bf s}_c)^T
487({\bf s} - {\bf s}_c) \nonumber \\
488  {\rm subject \ to } & \beta({\bf s}_c) + \nabla_s \beta({\bf s}_c)^T
489({\bf s} - {\bf s}_c) \ge \bar{\beta} \\
490& {\parallel {\bf s} - {\bf s}_c \parallel}_\infty \le \Delta^k \nonumber
491\label{eq:rbdo_surr}
492\end{eqnarray}
493and trust-region surrogate-based optimization for robust design
494employs surrogate models of $f$ and $\sigma^2$ within a trust region
495$\Delta^k$ centered at ${\bf s}_c$:
496\begin{eqnarray}
497  {\rm minimize }     & f({\bf s}_c) + \nabla_s f({\bf s}_c)^T
498({\bf s} - {\bf s}_c) \nonumber \\
499  {\rm subject \ to } & \sigma^2({\bf s}_c) + \nabla_s \sigma^2({\bf s}_c)^T
500({\bf s} - {\bf s}_c) \le \bar{\sigma}^2 \\
501& {\parallel {\bf s} - {\bf s}_c \parallel}_\infty \le \Delta^k \nonumber
502\label{eq:rdo_surr}
503\end{eqnarray}
504Second-order local surrogates may also be employed, where the Hessians
505are typically approximated from an accumulation of curvature
506information using quasi-Newton updates~\cite{Noc99} such as
507Broyden-Fletcher-Goldfarb-Shanno (BFGS, Eq.~\ref{eq:bfgs}) or
508symmetric rank one (SR1, Eq.~\ref{eq:sr1}).  The sequential approach
509is available for probabilistic expansions using PCE and SC.
510
511\subsubsection{Multifidelity SEBDO} \label{ouu:sebdo:form:mf}
512
513The multifidelity OUU approach is another trust-region surrogate-based
514approach.  Instead of the surrogate UQ model being a simple data fit
515(in particular, first-/second-order Taylor series model) of the truth
516UQ model results, distinct UQ models of differing fidelity are now
517employed.  This differing UQ fidelity could stem from the fidelity of
518the underlying simulation model, the fidelity of the UQ algorithm, or
519both.  In this section, we focus on the fidelity of the UQ algorithm.
520For reliability methods, this could entail varying fidelity in
521approximating assumptions (e.g., Mean Value for low fidelity, SORM for
522high fidelity), and for stochastic expansion methods, it could involve
523differences in selected levels of $p$ and $h$ refinement.
524
525%Here we will explore multifidelity stochastic models and employ
526%first-order additive corrections, where the meaning of multiple
527%fidelities is expanded to imply the quality of multiple UQ analyses,
528%not necessarily the fidelity of the underlying simulation model.  For
529%example, taking an example from the reliability method family, one
530%might employ the simple Mean Value method as a ``low fidelity'' UQ
531%model and take SORM as a ``high fidelity'' UQ model.  In this case,
532%the models do not differ in their ability to span a range of design
533%parameters; rather, they differ in their sets of approximating
534%assumptions about the characteristics of the response function.
535
536Here, we define UQ fidelity as point-wise accuracy in the design space
537and take the high fidelity truth model to be the probabilistic
538expansion PCE/SC model, with validity only at a single design point.
539The low fidelity model, whose validity over the design space will be
540adaptively controlled, will be either the combined expansion PCE/SC
541model, with validity over a range of design parameters, or the MVFOSM
542reliability method, with validity only at a single design point.  The
543combined expansion low fidelity approach will span the current trust
544region of the design space and will be reconstructed for each new
545trust region.  Trust region adaptation will ensure that the combined
546expansion approach remains sufficiently accurate for design purposes.
547By taking advantage of the design space spanning, one can eliminate
548the cost of multiple low fidelity UQ analyses within the trust region,
549with fallback to the greater accuracy and higher expense of the
550probabilistic expansion approach when needed.  The MVFOSM low fidelity
551approximation must be reformed for each change in design variables,
552but it only requires a single evaluation of a response function and
553its derivative to approximate the response mean and variance from the
554input mean and covariance
555(Eqs.~\ref{eq:mv_mean1}--\ref{eq:mv_std_dev}) from which
556forward/inverse CDF/CCDF reliability mappings can be generated using
557Eqs.~\ref{eq:mv_ria}--\ref{eq:mv_pma}.  This is the least expensive UQ
558option, but its limited accuracy\footnote{MVFOSM is exact for linear
559  functions with Gaussian inputs, but quickly degrades for nonlinear
560  and/or non-Gaussian.} may dictate the use of small trust regions,
561resulting in greater iterations to convergence.  The expense of
562optimizing a combined expansion, on the other hand, is not
563significantly less than that of optimizing the high fidelity UQ model,
564but its representation of global trends should allow the use of larger
565trust regions, resulting in reduced iterations to convergence.
566%While conceptually different, in the end, this approach is
567%similar to the use of a global data fit surrogate-based optimization
568%at the top level in combination with the probabilistic expansion PCE/SC
569%at the lower level, with the distinction that the multifidelity approach
570%embeds the design space spanning within a modified PCE/SC process
571%whereas the data fit approach performs the design space spanning
572%outside of the UQ (using data from a single unmodified PCE/SC process,
573%which may now remain zeroth-order).
574The design derivatives of each of the PCE/SC expansion models provide
575the necessary data to correct the low fidelity model to first-order
576consistency with the high fidelity model at the center of each trust
577region, ensuring convergence of the multifidelity optimization process
578to the high fidelity optimum.  Design derivatives of the MVFOSM
579statistics are currently evaluated numerically using forward finite
580differences.
581
582Multifidelity optimization for reliability-based design can be
583formulated as:
584\begin{eqnarray}
585  {\rm minimize }     & f({\bf s}) \nonumber \\
586  {\rm subject \ to } & \hat{\beta_{hi}}({\bf s}) \ge \bar{\beta} \\
587& {\parallel {\bf s} - {\bf s}_c \parallel}_\infty \le \Delta^k \nonumber
588\label{eq:rbdo_mf}
589\end{eqnarray}
590and multifidelity optimization for robust design can be formulated as:
591\begin{eqnarray}
592  {\rm minimize }     & f({\bf s}) \nonumber \\
593  {\rm subject \ to } & \hat{\sigma_{hi}}^2({\bf s}) \le \bar{\sigma}^2 \\
594& {\parallel {\bf s} - {\bf s}_c \parallel}_\infty \le \Delta^k \nonumber
595\label{eq:rdo_mf}
596\end{eqnarray}
597where the deterministic objective function is not approximated and
598$\hat{\beta_{hi}}$ and $\hat{\sigma_{hi}}^2$ are the approximated
599high-fidelity UQ results resulting from correction of the low-fidelity
600UQ results.  In the case of an additive correction function:
601\begin{eqnarray}
602\hat{\beta_{hi}}({\bf s})    &=& \beta_{lo}({\bf s}) +
603\alpha_{\beta}({\bf s})  \label{eq:corr_lf_beta} \\
604\hat{\sigma_{hi}}^2({\bf s}) &=& \sigma_{lo}^2({\bf s}) +
605\alpha_{\sigma^2}({\bf s}) \label{eq:corr_lf_sigma}
606\end{eqnarray}
607where correction functions $\alpha({\bf s})$ enforcing first-order
608%and quasi-second-order
609consistency~\cite{Eld04} are typically employed.  Quasi-second-order
610correction functions~\cite{Eld04} can also be employed, but care
611must be taken due to the different rates of curvature accumulation
612between the low and high fidelity models.  In particular, since the
613low fidelity model is evaluated more frequently than the high fidelity
614model, it accumulates curvature information more quickly, such that
615enforcing quasi-second-order consistency with the high fidelity model
616can be detrimental in the initial iterations of the
617algorithm\footnote{Analytic and numerical Hessians, when
618available, are instantaneous with no accumulation rate concerns.}.
619Instead, this consistency should only be enforced when sufficient high
620fidelity curvature information has been accumulated (e.g., after $n$
621rank one updates).
622
623
624
625\section{Sampling-based OUU}\label{ouu:sampling}
626
627Gradient-based OUU can also be performed using random sampling methods.
628In this case, the sample-average approximation to the design derivative
629of the mean and standard deviation are:
630\begin{eqnarray}
631  \frac{d\mu}{ds}    &=& \frac{1}{N} \sum_{i=1}^N \frac{dQ}{ds} \\
632  \frac{d\sigma}{ds} &=& \left[ \sum_{i=1}^N (Q \frac{dQ}{ds})
633    - N \mu \frac{d\mu}{ds} \right] / (\sigma (N-1))
634\end{eqnarray}
635This enables design sensitivities for mean,  standard deviation or
636variance (based on {\tt final\_moments} type), and forward/inverse
637reliability index mappings
638($\bar{z} \rightarrow \beta$, $\bar{\beta} \rightarrow z$).
639
640% TO DO: Multilevel MC ...
641