Next Article in Journal
Illustrative Application of the nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems to the Nordheim–Fuchs Reactor Dynamics/Safety Model
Previous Article in Journal
The Contribution of Small Modular Reactors to the Resilience of Power Supply
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N): Mathematical Framework

by
Dan Gabriel Cacuci
Center for Nuclear Science and Energy, University of South Carolina, Columbia, SC 29208, USA
J. Nucl. Eng. 2022, 3(3), 163-190; https://doi.org/10.3390/jne3030010
Submission received: 27 April 2022 / Revised: 9 June 2022 / Accepted: 15 June 2022 / Published: 21 June 2022

Abstract

:
This work presents the nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N), which enables the most efficient computation of exactly determined expressions of arbitrarily high-order sensitivities of generic nonlinear system responses with respect to model parameters, uncertain boundaries, and internal interfaces in the model’s phase space. The mathematical framework underlying the nth-CASAM-N is proven to be correct by using mathematical induction. The nth-CASAM-N is formulated in linearly increasing higher-dimensional Hilbert spaces—as opposed to exponentially increasing parameter-dimensional spaces—thus overcoming the curse of dimensionality in sensitivity analysis of nonlinear systems.

1. Introduction

The high-order sensitivities (i.e., functional derivatives) of model responses (i.e., results of interest produced by models) with respect to model parameters are notoriously difficult to compute for large-scale models involving many parameters. Furthermore, the computation of higher-order sensitivities by conventional methods is subject to the curse of dimensionality, a term coined by Belmann [1] to describe phenomena in which the number of computations increases exponentially in the respective phase space. In the particular case of sensitivity analysis, the number of large-scale computations increases exponentially in the parameter phase space as the order of sensitivities increases. It is known that the adjoint method of sensitivity analysis is the most efficient method for computing exactly first-order sensitivities. The idea underlying the computation of response sensitivities with respect to model parameters using adjoint operators was first used by Wigner [2] to analyze first-order perturbations in linear problems of interest to nuclear reactor physics. Cacuci [3,4] is credited (see, e.g., [5,6]) for having conceived the rigorous first-order adjoint sensitivity analysis methodology for generic large-scale nonlinear (as opposed to linearized) systems involving generic operator responses and having introduced these principles to the earth, atmospheric, and other sciences.
The general mathematical framework for the second-order adjoint sensitivity analysis methodology for linear and nonlinear systems, respectively, was conceived by Cacuci [7,8]. The unparalleled efficiency of the second-order adjoint sensitivity analysis methodology for linear systems was demonstrated [9] by applying this methodology to compute exactly the 21,976 first-order sensitivities and 482,944,576 second-order sensitivities (of which 241,483,276 are distinct from each other) for an OECD/NEA reactor physics benchmark [10], which is representative of a large-scale system that involves many (21,976, in this illustrative example) parameters. This paradigm large-scale system cannot be analyzed comprehensively by any other (including statistical) methods. The results obtained in [9] indicated that many second-order sensitivities were much larger than the largest first-order ones (contrary to the widely held belief to the contrary, for reactor physics systems), which motivated the investigation [11] of the largest third-order and, subsequently, fourth-order sensitivities. The evident conclusions that resulted from these works [9,11] are that the consideration of only the first-order sensitivities is insufficient for making credible predictions regarding the expected values and uncertainties (variances, covariances, skewness) of calculated and predicted/adjusted responses. At the very least, the second-order sensitivities must also be computed in order to enable the quantitative assessment of their impact on the predicted model responses. Consequently, Cacuci [12] has conceived the mathematical framework of the nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Response-Coupled Forward/Adjoint Linear Systems (abbreviated as “nth-CASAM-L”). The nth-CASAM-L enables the efficient computation of exactly determined expressions of arbitrarily high-order sensitivities of a generic system response—which can depend on both the forward and adjoint state functions—with respect to all of the parameters that characterize the physical system. The qualifier “comprehensive” is employed in order to highlight that the model parameters considered within the framework of the nth-CASAM-L include the system’s uncertain boundaries and internal interfaces in phase space. The nth-CASAM-L mathematical framework was developed specifically for linear systems because the most important model responses produced by such systems are various Lagrangian functionals, which depend simultaneously on both the forward and adjoint state functions governing the respective linear system. Included among such functionals are the Raleigh quotient for computing eigenvalues and/or separation constants when solving linear partial differential equations, and the Schwinger and Rousopoulos functionals (see, e.g., [13,14,15,16]), which play a fundamental role in optimization and control procedures, and the derivation of numerical methods for solving equations (differential, integral, integro-differential). Responses that depend on both the forward and adjoint state functions underlying the respective model can occur only for linear systems, because responses in nonlinear systems can only depend on the system’s forward state functions (since nonlinear operators do not admit adjoint operators). Consequently, the sensitivity analysis of responses that simultaneously involve both forward and adjoint state functions makes it necessary to treat linear models/systems in their own right, rather than treating them as particular cases of nonlinear systems.
The nth-CASAM-L provides the tools for the exact and efficient computation of higher-order sensitivities for response-coupled forward/adjoint linear systems while overcoming the curse of dimensionality that has hindered their computation thus far. In parallel, Cacuci [17,18] has also developed the 5th-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (5th-CASAM-N), thereby generalizing the original mathematical framework [3,4] for nonlinear systems. The 5th-CASAM-N enables the efficient and exact computation of response sensitivities, up to the fifth order, with respect to imprecisely known (i.e., uncertain) model boundaries and/or internal interfaces. The material to be presented in this work generalizes the 5th-CASAM-N to enable the exact and efficient computation of sensitivities of arbitrarily high order. This general methodology is called the nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N) and is presented in Section 2 of this work. The nth-CASAM-N is formulated in linearly increasing higher-dimensional Hilbert spaces (as opposed to exponentially increasing parameter-dimensional spaces), thus overcoming the curse of dimensionality in sensitivity analysis of nonlinear systems, enabling the most efficient computation of exactly determined expressions of arbitrarily high-order sensitivities of generic nonlinear system responses with respect to model parameters, uncertain boundaries, and internal interfaces in the model’s phase space. The mathematical framework underlying the nth-CASAM-N is proven to be correct by using mathematical induction. Section 3 concludes this work by discussing the significance of the nth-CASAM-N.

2. The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N)

The general framework of the nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Response-Coupled Forward/Adjoint Linear Systems (nth-CASAM-L) is constructed based on the pattern that emerges from the 4th-CASAM-N [17] and the 5th-CASAM-N [18]. The generic mathematical model of a nonlinear system, which comprises uncertain (i.e., imprecisely known) model parameters, boundaries, and interfaces, is detailed in Appendix A, since it is the same as that used in [17,18]. The validity of mathematical methodology underlying the nth-CASAM-L is established in this section by using a proof by mathematical induction comprising the usual steps, as follows:
  • Establish the general pattern underlying the nth-CASAM-N for an arbitrarily high-order n .
  • Prove that the general pattern underlying the nth-CASAM-N is valid for the lowest values of n , i.e., n = 1 and n = 2 .
  • Assuming that that the pattern is valid for an arbitrarily high-order n , prove that the pattern is also valid for n + 1 .

2.1. The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N)

The nth-order sensitivity R n j n ; ; j 1 ; U n ; A ( n ) ; α n R u x ; α / α j n α j 1 of the response R u x ; α with respect to the parameters α j 1 , , α j n is obtained, for each index j 1 , , j n = 1 , , T P , from the expression of the total first-order G-differential of each of the (n − 1)th-order sensitivities, each of which is expected to have the following expression for j 1 , , j n 1 = 1 , , T P :
R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α n 1 R u x ; α / α j n 1 α j 1 λ 1 α ω 1 α λ T I α ω T I α S n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α d x 1 d x T I .
Generalizing the pattern established in Appendix B for the nth-CASAM-N for n = 1 , 2 , 3 , the total first-order G-differential of the (n − 1)th-order sensitivity R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α is expected to have the following expression:
δ R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α α 0 j n = 1 T P R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α α j n α 0 δ α j n + δ R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α ; V n 1 ; δ A ( n 1 ) i n d
where the quantity δ R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α ; V n 1 ; δ A ( n 1 ) i n d denotes the indirect-effect term, which is defined as follows:
δ R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; V n 1 ; δ A ( n 1 ) ; α i n d λ 1 α ω 1 α λ T I α ω T I α S n 1 U n 1 x V n 1 2 n 2 ; x + S n 1 A ( n 1 ) x δ A ( n 1 ) 2 n 2 ; x α 0 d x 1 d x T I .
The vectors V n 1 2 n 2 ; j n 3 ; ; j 1 ; x δ U n 1 2 n 2 ; j n 3 ; ; j 1 ; x and δ A ( n 1 ) 2 n 2 ; j n 2 ; ; j 1 ; x are the solution of the following nth-Level Variational Sensitivity System (nth-LVSS), which is obtained by concatenating the (n − 1)th-LVSS together with the G-differentiated (n − 1)th-LASS for j 1 , , j n 1 = 1 , , T P :
V M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α V n 2 n 1 ; j n 2 ; ; j 1 ; x α 0 = Q V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α α 0 , x Ω x ,
B V n 2 n 1 ; U n 2 n 1 ; x ; V n 2 n 1 ; x ; α ; δ α α 0 = 0 2 n 1 ; x Ω x α 0 ,
The variational matrix V M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α comprises 2 n 1 × 2 n 1 block-matrices, each comprising T D 2 components/elements; thus, the matrix V M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α comprises a total of 2 n 1 × 2 n 1 T D 2 components/elements. Each of the vectors V n 2 n 1 ; j n 2 ; ; j 1 ; x , Q V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α , and B V n 2 n 1 ; U n 2 n 1 ; x ; V 4 2 n 1 ; x ; α ; δ α comprises 2 n 1   T D -dimensional vectors, so that each of these vectors comprises 2 n 1 × T D components/elements. The quantity 0 2 n 1 denotes a block-vector with 2 n 1 components, with each component being a T D -dimensional vector with identically zero components. The various quantities appearing in Equations (4) and (5) are defined as follows:
V M n 2 n 1 × 2 n 1 ; x ; α V M n 1 2 n 2 × 2 n 2 ; x 0 2 n 2 × 2 n 2 V M 21 n 2 n 2 × 2 n 2 ; x V M 22 n 2 n 2 × 2 n 2 ; x ;
U n 2 n 1 ; j n 2 ; ; j 1 ; x U n 1 2 n 2 ; j n 3 ; ; j 1 ; x A ( n 1 ) 2 n 2 ; j n 2 ; ; j 1 ; x ;
V n 2 n 1 ; j n 2 ; ; j 1 ; x δ U n 2 n 1 ; j n 2 ; ; j 1 ; x = V n 1 2 n 2 ; j n 3 ; ; j 1 ; x δ A ( n 1 ) 2 n 2 ; j n 2 ; ; j 1 ; x = v 1 x , δ a 1 x , δ a 2 1 ; j 1 ; x , δ a 2 2 ; j 1 ; x , , δ a ( n 1 ) 1 ; j n 2 ; ; j 1 ; x , , δ a ( n 1 ) 2 n 2 ; j n 2 ; ; j 1 ; x ;
V M 21 n 2 n 2 × 2 n 2 ; j n 2 ; ; j 1 ; x Q A n 1 2 n 2 ; j n 2 ; ; j 1 ; U n 2 2 n 2 ; j n 3 ; ; j 1 ; x ; U n 1 2 n 2 ; j n 3 ; ; j 1 ; x + A M n 1 2 n 2 × 2 n 2 ; U n 1 2 n 2 ; x ; α A ( n 1 ) 2 n 2 ; j n 2 ; ; j 1 ; x U n 1 2 n 2 ; j n 3 ; ; j 1 ; x ;
V M 22 n 2 n 2 × 2 n 2 ; x A M n 1 2 n 2 × 2 n 2 ; U n 1 2 n 2 ; j n 3 ; ; j 1 ; x ; α ;
Q V n 2 n 1 ; j 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α Q V n 2 2 n 2 ; U n 1 2 n 2 ; x ; α ; δ α Q 2 n 2 n 2 ; U n 2 n 1 ; x ; α ; δ α q V n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α , , q V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α ;
q V n i ; U n x ; α ; δ α j n = 1 T P s V n i ; j n 2 ; ; j 1 ; U n x ; α δ α j n ; i = 1 , ; 2 n 1 ;
Q 2 n 2 n 2 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α Q A n 1 2 n 2 ; j 1 ; U n 1 x ; α α α A M n 1 2 n 2 × 2 n 2 ; U n 1 2 n 2 ; x ; α A ( n 1 ) 2 n 2 ; x α α ;
B V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; V n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α B V n 1 2 n 2 ; U n 1 2 n 2 ; j n 3 ; ; j 1 ; x ; V n 1 2 n 2 ; j n 3 ; ; j 1 ; x ; α ; δ α δ B A n 1 2 n 2 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; V n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α .
Solving the nth-LVSS would require O T P n large-scale computations, which is unrealistic for large-scale systems comprising many parameters. The nth-CASAM-N circumvents the need to solve the nth-LVSS by deriving an alternative expression for the indirect-effect term defined in Equation (3), in which the function V n 2 n 1 ; j n 2 ; ; j 1 ; x is replaced by a nth-level adjoint function, denoted as A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x a ( n ) 1 ; j n 1 ; ; j 1 ; x , , a ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x H n Ω x , which is independent of parameter variations. The elements of the Hilbert space H n Ω x are block-vectors of the form Ψ ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x ψ ( n ) 1 ; j n 1 ; ; j 1 ; x , , ψ ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x , comprising 2 n 1   T D -dimensional vectors of the form ψ n i ; x ψ 1 n i ; x , , ψ T D n i ; x H 1 Ω x , i = 1 , , 2 n . The inner product of two vectors Ψ ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x and Φ ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x in the Hilbert space H n Ω x is denoted as Ψ ( n ) 2 n 1 ; x , Φ ( n ) 2 n 1 ; x n and defined as follows:
Ψ ( n ) 2 n 1 ; x , Φ ( n ) 2 n 1 ; x n i = 1 2 n 1 ψ ( n ) i ; x , φ ( n ) i ; x 1 .
The nth-Level Adjoint Sensitivity System (nth-LASS) for the nth-level adjoint function A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x is obtained by using the inner product defined in Equation (15) as follows:
A ( n ) 2 n 1 ; x , V M n 2 n 1 ; x 2 α 0 = P n U n ; A ( n ) ; V n ; α Ω x α 0 + V n 2 n 1 ; x , A M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; x ; α A ( n ) 2 n 1 ; x n α 0 ,
where the quantity P n U n ; A ( n ) ; V n ; α Ω x α 0 denotes the corresponding bilinear concomitant on the domain’s boundary, evaluated at the nominal values for the parameters and respective state functions.
In terms of the nth-level adjoint function A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x , the indirect-effect term defined by Equation (3) will have the following expression:
δ R n 1 j n 1 ; ; j 1 ; U n 1 ; A ( n 1 ) ; α ; V n 1 ; δ A ( n 1 ) i n d = P ^ n U n ; A ( n ) ; δ α Ω x α 0 + A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x , Q V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α n α 0 .
where A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x is the solution of the following nth-Level Adjoint Sensitivity System (nth-LASS):
A M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; x ; α A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x = Q A n 2 n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ,
subject to boundary conditions represented in operator form as follows:
B A n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x ; α α 0 = 0 2 n 1 , x Ω x α 0 ; j 1 = 1 , , T P ; j 2 = 1 , , j 1 ; ; j n 1 = 1 , , j n 2 .
In Equation (17), the quantity P ^ n U n ; A ( n ) ; δ α Ω x α 0 denotes residual boundary terms that may have not vanished automatically after having used the boundary conditions provided in Equations (5) and (19) to eliminate from Equation (17) all unknown values of the nth-level variational function V n 2 n 1 ; j n 2 ; ; j 1 ; x .
The quantities that appear in the definition of the nth-LASS, cf. Equations (18) and (19), are defined as follows:
AM n 2 n 1 × 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; AM ; α VM n 2 n 1 × 2 n 1 ; U n ; α * = VM n 1 2 n 2 × 2 n 2 * VM 21 n 2 n 2 × 2 n 2 * 0 2 n 2 × 2 n 2 VM 22 n 2 n 2 × 2 n 2 * ,
Q A n 2 n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; x ; α q A n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; x ; α , , q A n 2 n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; x ; α , j 1 = 1 , , T P ; ; j n 1 = 1 , , j n 2 ;
q A n 1 ; j n 1 ; ; j 1 ; U n x ; α S n 1 j n 1 ; ; j 1 ; U n ; α / u x ;
q A n 2 ; j n 1 ; ; j 1 ; U n x ; α S n 1 j n 1 ; ; j 1 ; U n ; α / a 1 x ;
f o r n 3 : q A n 2 k + i ; j n 1 ; ; j 1 ; U n x ; α S n 1 j n 1 ; ; j 1 ; U n ; α a ( k + 1 ) i ; j k ; ; j 1 ; x ; k = 1 , , n 2 ; i = 1 , , 2 k .
The final expression of the total differential expressed by Equation (2) is obtained by inserting into Equation (2) the expression for the indirect-effect term obtained in Equation (17), which yields the following expression for the nth-order sensitivities R n j n ; ; j 1 ; U n ; A ( n ) ; α n R u x ; α / α j n α j 1 of the response R u x ; α with respect to the parameters α j 1 , , α j n , for j 1 = 1 , , T P ; ; j n 1 = 1 , , j n 2 :
R n j n ; ; j 1 ; U n ; A n ; α n R u x ; α / α j n α j 1 = R n 1 j n 1 ; ; j 1 ; U n 1 x ; A n 1 x ; α α j n P ^ n U n ; A n ; δ α Ω x α j n + i = 1 2 n 1 a n i ; j n 1 ; ; j 1 ; x , s V n i ; j n ; ; j 1 ; U n x ; α 1 . λ 1 α ω 1 α λ T I α ω T I α S n j n ; ; j 1 ; U n 2 n 1 ; x ; A n 2 n 1 ; x ; α d x 1 d x T I .

2.2. The nth-CASAM-N for n = 1 and n = 2

Setting n = 1 in the mathematical framework of the nth-CASAM-N conjectured above in Section 2.1 yields the following expressions for the particular case n = 1 :
(i)
Expression of the model response
R 0 U 0 ; α R u x ; α λ 1 α ω 1 α λ T I α ω T I α S u x ; α d x 1 d x T I ,
(ii)
Expression of the first-order response sensitivities for j 1 = 1 , , T P :
R 1 j 1 ; U 1 2 0 ; x ; A ( 1 ) 2 0 ; x ; α = R u x ; α α j 1 = P ^ 1 U 1 ; A ( 1 ) ; δ α Ω x α j 1 + a ( 1 ) x , s V 1 U 1 x ; α 1 .
The various quantities that appear in Equation (27) take on the following particular forms for n = 1 :
U 1 2 0 ; x u x ; A ( 1 ) 2 0 ; x a ( 1 ) x ; s V 1 j 1 ; u ; α Q α N u ; α α j 1 .
The first-level adjoint sensitivity function A 1 2 0 ; x a 1 x is the solution of the following 1st-LASS obtained by setting n = 1 in Equations (18) and (19):
A M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α A 1 2 0 ; x α 0 = Q A 1 2 0 ; U 1 2 0 ; x ; α ; δ α α 0 , x Ω x ,
B A 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α α 0 = 0 , x Ω x α 0 .
where
A M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α N 1 u ; α * ;
Q A 1 2 0 ; U 1 2 0 ; x ; α ; δ α q A 1 u x ; α S u ; α / u α 0 ;
B V 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α b A 1 u ; a 1 ; α .
Comparing the expressions obtained above in Equations (26)–(33) with the expressions obtained in Appendix B.3 reveals that the corresponding expressions are identical to each other, which proves the correctness for the particular case n = 1 of the conjectured general expressions underlying the nth-CASAM-N.
Setting n = 2 in the mathematical framework of the nth-CASAM-N conjectured above in Section 2.1 yields the following expressions for the second-order response sensitivities for j 1 , j 2 = 1 , , T P :
R 2 j 2 ; j 1 ; U 2 2 ; x ; A 2 2 ; j 1 ; x ; α 2 R u x ; α α j 2 α j 1 = R 1 j 1 ; u x ; a 1 x ; α α j 2 P ^ 2 U 2 2 ; x ; A 2 2 ; j 1 ; x ; α Ω x α j 2 + i = 1 2 a 2 i ; j 1 ; x , s V 2 i ; j 2 ; U 2 2 ; x ; α 1 ,
where A ( 2 ) 2 ; j 1 ; x a ( 2 ) 1 ; j 1 ; x , a ( 2 ) 2 ; j 1 ; x H 2 Ω x is the solution of the following 2nd-Level Adjoint Sensitivity System (2nd-LASS) for each j 1 = 1 , , T P :
A M 2 2 × 2 ; U 2 2 ; x ; α A ( 2 ) 2 ; j 1 ; x α 0 = Q A 2 2 ; j 1 ; U 2 2 ; x ; α α 0 , j 1 = 1 , , T P ; x Ω x ,
B A 2 2 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α α 0 = 0 2 ; j 1 = 1 , , T P ; x Ω x α 0 .
Q A 2 2 ; j 1 ; U 2 2 ; x ; α q A 2 1 ; j 1 ; U 2 ; α q A 2 2 ; j 1 ; U 2 ; α S 1 j 1 ; u x ; a 1 x ; α ; v 1 x / u S 1 j 1 ; u x ; a 1 x ; α ; v 1 x / a 1 , j 1 = 1 , , T P .
Comparing the expressions obtained in Equations (26)–(33) with the expressions obtained in Appendix B.3 reveals that the corresponding expressions are identical to each other, which proves the correctness for the particular case n = 2 of the conjectured general expressions underlying the nth-CASAM-N.

2.3. The (n + 1)th- CASAM-N

The nth-order sensitivities R n j n ; ; j 1 ; U n ; A ( n ) ; α n R u x ; α / α j n α j 1 R 4 j 4 ; j 3 ; j 2 ; j 1 ; U 4 ; A ( 4 ) ; α will be assumed to satisfy the conditions stated in Equations (A10) and (A11) in Appendix B for each j 1 , , j n = 1 , , T P . Hence, the first-order total G-differential of R n j n ; ; j 1 ; U n ; A ( n ) ; α will exist and will be linear in the variations V n 2 n 1 ; j n 2 ; ; j 1 ; x δ U n 1 2 n 1 ; j n 2 ; ; j 1 ; x and δ A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x in a neighborhood around the nominal values of the parameters and the respective state functions. By definition, the first-order total G-differential of R n j n ; ; j 1 ; U n ; A ( n ) ; α is given by the following expression:
δ R n j n ; ; j 1 ; U n ; A n ; α α 0 d d ε R n j n ; ; j 1 ; U n + ε V n ; A n + ε δ A n ; α + ε δ α ε = 0 j n + 1 = 1 T P R n ; U n ; A n ; α α j n + 1 α 0 δ α j n + 1 + δ R n j n ; ; j 1 ; U n ; A n ; α ; V n x ; δ A n x i n d ,
where the quantity δ R n j n ; ; j 1 ; U n ; A ( n ) ; α ; V n ; δ A ( n ) i n d denotes the indirect-effect term, which is defined as follows:
δ R n j n ; ; j 1 ; U n ; A ( n ) ; α ; V n ; δ A ( n ) i n d λ 1 α ω 1 α λ T I α ω T I α S n U n x V n 2 n 1 ; x + S n A ( n ) x δ A ( n ) 2 n 1 ; x α 0 d x 1 d x T I
The vectors V n 2 n 1 ; j n 2 ; ; j 1 ; x δ U n 2 n 1 ; j n 2 ; ; j 1 ; x and δ A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x , which are needed in order to evaluate the indirect-effect term δ R n j n ; ; j 1 ; U n ; A ( n ) ; α ; V n ; δ A ( n ) i n d , are the solutions of the following (n + 1)th-Level Variational Sensitivity System, which is obtained by concatenating the nth-LVSS defined by Equations (4) and (5) together with the G-differentiated nth-LASS for j 1 , , j n = 1 , , T P :
V M n + 1 2 n × 2 n ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α V n + 1 2 n ; j n 1 ; ; j 1 ; x α 0 = Q V n + 1 2 n ; U n 2 n ; j n 2 ; ; j 1 ; x ; α ; δ α α 0 , x Ω x ,
B V n + 1 2 n ; U n + 1 2 n ; x ; V n + 1 2 n ; x ; α ; δ α α 0 = 0 2 n ; x Ω x α 0 ,
where
U n + 1 2 n ; j n 1 ; ; j 1 ; x U n 2 n 1 ; j n 2 ; ; j 1 ; x A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x ;
V n + 1 2 n ; j n 1 ; ; j 1 ; x δ U n + 1 2 n ; j n 1 ; ; j 1 ; x = V n 2 n 1 ; j n 2 ; ; j 1 ; x δ A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x = v 1 x , δ a 1 x , δ a 2 1 ; j 1 ; x , δ a 2 2 ; j 1 ; x , , δ a ( n ) 1 ; j n 1 ; ; j 1 ; x , , δ a ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x ;
V M n + 1 2 n × 2 n ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α V M n 2 n 1 × 2 n 1 ; x ; α 0 2 n 1 × 2 n 1 V M 21 n + 1 2 n 1 × 2 n 1 ; x ; α V M 22 n + 1 2 n 1 × 2 n 1 ; x ; α ;
V M 21 n + 1 2 n 1 × 2 n 1 ; j n 1 ; ; j 1 ; x ; α Q A n 2 n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α U n 2 n 1 ; j n 2 ; ; j 1 ; x + A M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; x ; α A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x U n 2 n 1 ; j n 2 ; ; j 1 ; x ;
V M 22 n + 1 2 n 1 × 2 n 1 ; x ; α A M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ;
Q V n + 1 2 n ; j 1 ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α Q V n 2 n 1 ; U n 2 n 1 ; x ; α ; δ α Q 2 n + 1 2 n 1 ; U n + 1 2 n ; x ; α ; δ α q V n + 1 1 ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α , , q V n + 1 2 n ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α ;
q V n i ; U n x ; α ; δ α j n = 1 T P s V n i ; j n 2 ; ; j 1 ; U n x ; α δ α j n ; i = 1 , ; 2 n 1 ;
Q 2 n + 1 2 n 1 ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α Q A n 2 n 1 ; j n 1 ; ; j 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; α α α A M n 2 n 1 × 2 n 1 ; U n 2 n 1 ; x ; α A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x α α ;
B V n + 1 2 n ; U n + 1 2 n ; j n 1 ; j 1 ; x ; V n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α B V n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; V n 2 n 1 ; j n 2 ; ; j 1 ; x ; α ; δ α δ B A n 2 n 1 ; U n 2 n 1 ; j n 2 ; ; j 1 ; x ; A n 2 n 1 ; j n 1 ; ; j 1 ; x ; α .
Solving the (n + 1)th-LVSS would require O T P n + 1 large-scale computations, which is unrealistic for large-scale systems comprising many parameters. The (n + 1)th-CASAM-N circumvents the need to solve the (n + 1)th-LVSS by deriving an alternative expression for the indirect-effect term defined in Equation (39), in which the function V n + 1 2 n ; j n 1 ; ; j 1 ; x is replaced by a (n + 1)th-level adjoint function, denoted as A ( n + 1 ) 2 n ; j n ; ; j 1 ; x a ( n + 1 ) 1 ; j n ; ; j 1 ; x , , a ( n + 1 ) 2 n ; j n ; ; j 1 ; x H n + 1 Ω x , which is independent of parameter variations. The elements of the Hilbert space H n + 1 Ω x are block-vectors of the form Ψ ( n + 1 ) 2 n ; j n ; ; j 1 ; x ψ ( n + 1 ) 1 ; j n ; ; j 1 ; x , , ψ ( n + 1 ) 2 n ; j n ; ; j 1 ; x , comprising 2 n   T D -dimensional vectors of the form ψ n + 1 i ; x ψ 1 n + 1 i ; x , , ψ T D n + 1 i ; x H 1 Ω x , i = 1 , , 2 n + 1 . The inner product of two vectors Ψ ( n + 1 ) 2 n ; j n ; ; j 1 ; x and Φ ( n + 1 ) 2 n ; j n ; ; j 1 ; x in the Hilbert space H n + 1 Ω x is denoted as Ψ ( n + 1 ) 2 n ; x , Φ ( n + 1 ) 2 n ; x n + 1 and defined as follows:
Ψ ( n + 1 ) 2 n ; x , Φ ( n + 1 ) 2 n ; x n + 1 i = 1 2 n ψ ( n + 1 ) i ; x , φ ( n + 1 ) i ; x 1 .
The (n + 1)th-Level Adjoint Sensitivity System [(n + 1)th-LASS ] for the (n + 1)th-level adjoint function A ( n + 1 ) 2 n ; j n ; ; j 1 ; x is obtained by using the inner product defined in Equation (51) as follows:
A ( n + 1 ) 2 n ; x , V M n + 1 2 n ; x 2 α 0 = P n + 1 U n + 1 ; A ( n + 1 ) ; V n + 1 ; α Ω x α 0 + V n + 1 2 n ; x , A M n + 1 2 n × 2 n ; U n + 1 2 n ; x ; α A ( n + 1 ) 2 n ; x n α 0 ,
where the quantity P n + 1 U n + 1 ; A ( n + 1 ) ; V n + 1 ; α Ω x α 0 denotes the corresponding bilinear concomitant on the domain’s boundary, evaluated at the nominal values for the parameters and respective state functions.
In terms of the (n + 1)th-level adjoint function A ( n + 1 ) 2 n ; j n ; ; j 1 ; x , the indirect-effect term defined by Equation (39) will have the following expression:
δ R n j n ; ; j 1 ; U n ; A ( n ) ; α ; V n ; δ A ( n ) i n d = P ^ n + 1 U n + 1 ; A ( n + 1 ) ; δ α Ω x α 0 + A ( n + 1 ) 2 n ; j n ; ; j 1 ; x , Q V n + 1 2 n ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ; δ α n α 0 .
where A ( n + 1 ) 2 n ; j n ; ; j 1 ; x is the solution of the following (n + 1)th-LASS:
A M n + 1 2 n × 2 n ; U n + 1 2 n ; x ; α A ( n + 1 ) 2 n ; j n ; ; j 1 ; x = Q A n + 1 2 n ; j n ; ; j 1 ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α ,
subject to boundary conditions represented in operator form as follows:
B A n + 1 2 n ; U n 2 n ; j n 1 ; ; j 1 ; x ; A ( n ) 2 n ; j n ; ; j 1 ; x ; α α 0 = 0 2 n , x Ω x α 0 ; j 1 = 1 , , T P ; j 2 = 1 , , j 1 ; ; j n = 1 , , j n 1 .
In Equation (53), the quantity P ^ n + 1 U n + 1 ; A ( n + 1 ) ; δ α Ω x α 0 denotes residual boundary terms that may have not vanished automatically after having used the boundary conditions provided in Equations (41) and (55) to eliminate from Equation (52) all unknown values of the (n + 1)th-level variational function V n + 1 2 n ; j n 1 ; ; j 1 ; x .
The quantities that appear in the definition of the (n + 1)th-LASS, cf. Equations (54) and (55), are defined as follows:
AM n + 1 2 n × 2 n ; U n + 1 2 n ; j n 1 ; ; j 1 ; x ; α VM n + 1 2 n × 2 n ; U n + 1 ; α * = VM n 2 n 1 × 2 n 1 * VM 21 n + 1 2 n 1 × 2 n 1 * 0 2 n 1 × 2 n 1 VM 22 n + 1 2 n 1 × 2 n 1 * ,
Q A n + 1 2 n ; j n ; ; j 1 ; U n + 1 2 n ; x ; α q A n + 1 1 ; j n ; ; j 1 ; U n + 1 2 n ; x ; α , , q A n + 1 2 n ; j n ; ; j 1 ; U n + 1 2 n ; x ; α , j 1 = 1 , , T P ; ; j n = 1 , , j n 1 ;
q A n + 1 1 ; j n ; ; j 1 ; U n + 1 x ; α S n j n ; ; j 1 ; U n + 1 ; α / u x ;
q A n + 1 2 ; j n ; ; j 1 ; U n + 1 x ; α S n j n ; ; j 1 ; U n + 1 ; α / a 1 x ;
f o r n 3 : q A n + 1 2 k + i ; j n ; ; j 1 ; U n + 1 x ; α S n j n ; ; j 1 ; U n + 1 ; α a ( k + 1 ) i ; j k ; ; j 1 ; x ; k = 1 , , n 1 ; i = 1 , , 2 k .
The final expression of the total differential expressed by Equation (17) is obtained by inserting into Equation (38) the expression for the indirect-effect term obtained in Equation (53), which yields the following expression for the (n + 1)th-order sensitivities R n + 1 j n + 1 ; ; j 1 ; U n + 1 ; A ( n + 1 ) ; α n + 1 R u x ; α / α j n + 1 α j 1 of the response R u x ; α with respect to the parameters α j 1 , , α j n + 1 , for j 1 = 1 , , T P ; ; j n = 1 , , j n 1 :
R n + 1 j n + 1 ; ; j 1 ; U n + 1 ; A n + 1 ; α n + 1 R u x ; α / α j n + 1 α j 1 = R n j n ; ; j 1 ; U n x ; A n x ; α α j n + 1 P ^ n + 1 U n + 1 ; A n + 1 ; δ α Ω x α j n + 1 + i = 1 2 n a n + 1 i ; j n ; ; j 1 ; x , s V n + 1 i ; j n + 1 ; ; j 1 ; U n + 1 x ; α 1 . λ 1 α ω 1 α λ T I α ω T I α S n + 1 j n + 1 ; ; j 1 ; U n + 1 2 n ; x ; A n + 1 2 n ; x ; α d x 1 d x T I .
The expression obtained in Equation (61) is identical to the expression that would be obtained by replacing the index n with (n + 1) in the expression obtained in Equation (25), thus completing the proof, by mathematical induction, of the validity/correctness of the conjectured general expressions underlying the nth-CASAM-N.

3. Discussion

This work has presented the nth-Order Comprehensive Sensitivity Analysis Methodology for Nonlinear Systems (abbreviated as “nth-CASAM-N”), which enables the hitherto very difficult, if not intractable, exact computation of arbitrarily high-order response sensitivities with respect to uncertain model and boundary parameters. The mathematical framework underlying the nth-CASAM-N is proven to be correct by using mathematical induction. The nth-CASAM-N is formulated in linearly increasing higher-dimensional Hilbert spaces—as opposed to exponentially increasing parameter-dimensional spaces—thus overcoming the curse of dimensionality in sensitivity analysis of nonlinear systems. For very large and complex models, the nth-CASAM-N remains the only practical method for computing response sensitivities comprehensively and accurately.
The question of when to stop computing progressively higher-order sensitivities has been addressed by Cacuci [19] in conjunction with the question of convergence of the Taylor-series expansion of the response in terms of the uncertain model parameters, since this Taylor-series expansion is the fundamental premise for the expressions provided by the “propagation of errors” methodology for the cumulants of the model response distribution in the phase space of model parameters. The convergence of this Taylor-series, which depends on both the response sensitivities to parameters and the uncertainties associated with the parameter distribution, must be ensured. This can be done by ensuring that the combination of parameter uncertainties and response sensitivities is sufficiently small to fall inside the radius of convergence of the Taylor-series expansion. If the Taylor-series fails to converge, targeted experiments must be performed in order to reduce the largest sensitivities as well as the largest uncertainties (particularly standard deviations) that affect the most important parameters, by applying the principles of the BERRU-PM [20] methodology to obtain best-estimate parameter values with reduced uncertainties (model calibration).
Taken together, the nth-CASAM-N and nth-CASAM-L are the only practical methods for computing response sensitivities comprehensively and accurately for large and complex models, as has been illustrated by the application [9,11] of the nth-CASAM-L to the OECD/NEA reactor physics benchmark. Illustrative paradigm applications of the nth-CASAM-N are currently being finalized.

Funding

This research received no external funding.

Institutional Review Board Statement

No applicable.

Informed Consent Statement

No applicable.

Data Availability Statement

No applicable.

Conflicts of Interest

The author declares no conflict of interest.

Nomenclature

Symbols
A ( n + 1 ) 2 n ; j n ; ; j 1 ; x (n + 1)th-level adjoint block-vector function
A M n + 1 2 n × 2 n ; U n + 1 2 n ; x ; α (n + 1)th-level adjoint block-matrix operator
α α 1 , , α T P Vector of imprecisely known model parameters
B A n + 1 2 n ; U n 2 n ; x ; A ( n ) 2 n ; x ; α (n + 1)th-level block-vector containing adjoint boundary conditions
B V n + 1 2 n ; U n + 1 2 n ; x ; V n + 1 2 n ; x ; α ; δ α (n + 1)th-level block-vector containing variational boundary conditions
δ A ( n ) 2 n 1 ; j n 1 ; ; j 1 ; x nth-level adjoint variational block-vector function
δ R n j n ; ; j 1 ; U n ; A ( n ) ; α nth-order Gateaux variation of a model response
λ i α , ω i α , i = 1 , , T I Lower and, respectively, upper end point of the interval of definition of independent variable x i
nth-CASAM-Nnth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems
nth-CASAM-Lnth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Response-Coupled Forward/Adjoint Linear Systems
nth-LASSnth-Level Adjoint Sensitivity System
nth-LVSSnth-Level Variational Sensitivity System
P m i x i Spectral expansion functions
P n + 1 U n + 1 ; A ( n + 1 ) ; V n + 1 ; α Ω x (n + 1)th-level bilinear concomitant evaluated on the model’s boundary in the phase space of independent variables
Q A n + 1 2 n ; j n ; ; j 1 ; U n + 1 ; α (n + 1)th-level block vector of adjoint source terms
Q V n + 1 2 n ; U n ; α ; δ α (n + 1)th-level block vector of variational source terms
R u x ; α ; R u x ; α Vector-valued and, respectively, scalar-valued model response
R n + 1 j n + 1 ; ; j 1 ; U n + 1 ; A ( n + 1 ) ; α n + 1 R u x ; α / α j n + 1 α j 1 (n + 1)th-order sensitivity of a model response with respect to the model’s parameters
U n + 1 2 n ; j n 1 ; ; j 1 ; x (n + 1)th-level forward block-vector function
V n + 1 2 n ; j n 1 ; ; j 1 ; x δ U n + 1 2 n ; j n 1 ; ; j 1 ; x (n + 1)th-level variational block-vector function
V M n + 1 2 n × 2 n ; U n + 1 ; α (n + 1)th-level variational block-matrix operator
T D Total number of dependent variables in the model
T I Total number of independent (phase-space) variables in the model
T P Total number of parameters in the model
x x 1 , , x T I vector comprising the model’s independent variables
Ω α Domain of definition for a model’s independent variables
Ω α Boundary of Ω α

Appendix A. Mathematical Modeling of a Nonlinear System with Imprecisely Known Parameters and Domain Boundaries

The computational model of a physical system comprises equations that relate the system’s independent variables and parameters to the system’s state variables. The model parameters usually stem from processes that are external to the system under consideration and are seldom, if ever, known precisely. The known characteristics of the model parameters may include their nominal (expected/mean) values and, possibly, higher-order moments or cumulants (i.e., variance/covariances, skewness, kurtosis), which are usually determined from experimental data and/or processes external to the physical system under consideration. Occasionally, just the lower and the upper bounds may be known for some model parameters, expressed by inequality and/or equality constraints that delimit the ranges of the system’s parameters that are known. Without loss of generality, the imprecisely known model parameters can be considered to be real-valued scalar quantities. These model parameters will be denoted as α 1 ,…, α T P , where T P denotes the total number of imprecisely known parameters underlying the model under consideration. For subsequent developments, it is convenient to consider that these parameters are components of a vector of parameters, denoted as α α 1 , , α T P E α T P , where E α is also a normed linear space and where T P denotes the T P -dimensional subset of the set of real scalars. The components of the T P -dimensional column vector α T P are considered to include imprecisely known geometrical parameters that characterize the physical system’s boundaries in the phase space of the model’s independent variables. The nominal parameter values will be denoted as α 0 α 1 0 , , α i 0 , , α T P 0 ; the superscript “0” will be used throughout this work to denote nominal values. Matrices will be denoted using capital bold letters, whereas vectors will be denoted using either capital or lower-case bold letters. The symbol “ ” will be used to denote “is defined as” or “is by definition equal to.” Transposition will be indicated by a dagger superscript.
The generic nonlinear model is considered to comprise T I independent variables, which will be denoted as x i , i = 1 , , T I and that are considered to be components of a T I -dimensional column vector denoted as x x 1 , , x T I T I , where the sub/superscript “TI” denotes the total number of independent variables. The vector x T I of independent variables is considered to be defined on a phase-space domain that will be denoted as Ω α and that is defined as follows: Ω α λ i α x i ω i α ; i = 1 , , T I . The lower boundary point of an independent variable is denoted as λ i α and the corresponding upper boundary point is denoted as ω i α . The boundary Ω α is also considered to be imprecisely known since it may depend on both geometrical parameters and material properties. A typical example of boundaries that depend on both geometrical parameters and material properties are the boundaries facing vacuum in models based on diffusion theory, where conditions are imposed on the extrapolated boundary of the respective spatial domain. The extrapolated boundary depends both on the imprecisely known physical dimensions of the problem’s domain and also on the medium’s properties, such as atomic number densities and microscopic transport cross-sections. The boundary of Ω α , which will be denoted as Ω α , comprises the set of all of the endpoints λ i α , ω i α , i = 1 , , T I , of the respective intervals on which the components of x are defined, i.e., Ω α λ i α ω i α , i = 1 , , T I .
A nonlinear physical system can be generally modeled by means of coupled equations, which can be represented in operator form as follows:
N u x , α = Q x , α , x Ω x α .
The quantities that appear in Equation (A1) are defined as follows:
  • u x u 1 x , , u T D x is a T D -dimensional column vector of dependent variables; the abbreviation T D denotes “total number of dependent variables.” The functions u i x , i = 1 , , T D , denote the system’s dependent variables (also called “state functions”); u x E u , where E u is a normed linear space over the scalar field F of real numbers.
  • N u x ; α N 1 u ; α , , N T D u ; α denotes a T D -dimensional column vector. The components N i u ; α , i = 1 , , T D are operators (including differential, difference, integral, distributions, and/or finite or infinite matrices) acting (usually) nonlinearly on the dependent variables u x , the independent variables x , and the model parameters α . The mapping N u ; α is defined on the combined domains of the model’s parameters and state functions, i.e., N : D E E Q , where D = D u D α , D u E u , D α E α , E = E u E α .
  • Q x , α q 1 x ; α , , q T D x ; α is a T D -dimensional column vector that represents inhomogeneous source terms, which usually depend nonlinearly on the uncertain parameters α . The vector Q x , α is defined on a normed linear space denoted as E Q , i.e., Q E Q .
The equalities in this work are considered to hold in the weak (distributional) sense. The right sides of Equation (A1) and of other various equations to be derived in this work may contain generalized functions/functionals, particularly Dirac-distributions and derivatives thereof.
Boundary and/or initial conditions must also be provided if differential operators appear in Equation (A1). In operator form, these boundaries and/or initial conditions are represented as follows:
B u x ; α ; x C x , α = 0 , x Ω x α .
where the column vector 0 has T D components, all of which are identically zero, i.e.,
0 ζ 1 , , ζ T D ; ζ i 0 , i = 1 , , T D .
In Equation (A2), the components B i u ; α , i = 1 , , T D of B u ; α B 1 u ; α , , B T D u ; α are nonlinear operators in u x and α , which are defined on the boundary Ω x α of the model’s domain Ω x α . The components C i x ; α , i = 1 , , T D of C x , α C 1 x ; α , , C T D x ; α comprise inhomogeneous boundary sources that are nonlinear functions of α .
Solving Equations (A1) and (A2) at the nominal parameter values, denoted as α 0 α 1 0 , , α i 0 , , α T P 0 , provides the nominal solution u 0 x , i.e., the vectors u 0 x and α 0 satisfy the following equations:
N u 0 x ; α 0 = Q x , α 0 , x Ω x ,  
B u 0 x ; α 0 ; x - C x , α 0 = 0 , x Ω x α 0 .
The results computed using a mathematical model are customarily called “model responses” (or “system responses,” “objective functions,” or “indices of performance”). In general, a function-valued (i.e., operator-type) response R u x ; α can be represented by a spectral expansion in multidimensional orthogonal polynomials or Fourier series of the form:
R u x ; α = m 1 m T I c m 1 m T I u x ; α P m 1 x 1 P m 2 x 2 P m T I x T I ,
where the quantities P m i x i , i = 1 , , T I , denote the corresponding spectral functions (e.g., orthogonal polynomials or Fourier exponential/trigonometric functions) and where the spectral Fourier coefficients c m 1 m T I u x ; α are defined as follows:
c m 1 m T I u x ; α λ 1 α ω 1 α λ i α ω i α λ T I α ω T I α R u x ; α P m 1 x 1 P m T I x T I D x 1 D x T I .
The coefficients c m 1 m T I u x ; α can themselves be considered model responses since the spectral polynomials P m i x i are perfectly well known, whereas the expansion coefficients will contain all of the dependencies (directly or indirectly through the state functions) of the respective response on the imprecisely known model parameters. This way, the sensitivity analysis of an operator-valued response R u x ; α can be reduced to the sensitivity analysis of the scalar-valued responses c m 1 m T I u x ; α .
A measurement of a physical quantity that depends on the model’s state functions and parameters can be considered to be a response denoted as R p u x ; α , which is to be evaluated at x = x p α , where x p α x 1 p α , x k p α , , x T I p α denotes the location in phase space of the specific measurement point. Such a measurement (or measurement-like) response can be represented mathematically as follows:
R p u x ; α λ 1 α ω 1 α λ T I α ω T I α F u x ; α ; x δ x 1 x 1 p α δ x T I x T I p α D x 1 D x T I .
where the function   F u x ; α ; x denotes the mathematical dependence of the measurement device on the model’s dependent variable(s), and where the quantity δ x i x i p α denotes the Dirac-delta functional. The measurement’s location in phase space, x p α , may itself be afflicted by measurement (experimental) uncertainties. Hence, it is convenient to consider the components of x p α to be included among the components of the vector α of model parameters, even though x p α appears only in the definition of the response but does not appear in Equations (A1) and (A2), which mathematically define the physical model. Thus, the physical system is defined to comprise both the system’s computational model and the system’s response. In most cases, the coordinates x k p α , k = 1 , , T I , will be independent (albeit uncertain) model parameters, in which case x k p α / α n = 1 , i f α n x k p and x k p α / α n = 0 , i f α n x k p .
The representations shown in Equations (A6)–(A8) indicate that model responses can be fundamentally analyzed by considering the following generic integral representation:
R u x ; α λ 1 α ω 1 α λ T I α ω T I α S u x ; α ; x D x 1 D x T I ,
where S u x ; α is a suitably differentiable nonlinear function of u x and of α . It is important to note that the components of α include not only parameters that appear in the equations defining the computational model per se, i.e., in Equations (A1) and (A2), but also include parameters that specifically occur only in the definition of the response under consideration.
It is also important to note that the system’s definition domain, Ω α , in phase space is considered to be imprecisely known and subject to uncertainties in the components of the vector of model parameters α . Therefore, the system domain’s boundary, Ω α , as well as the model response R u x ; α , will be affected by the boundary uncertainties that affect the endpoints λ i α , ω i α , i = 1 , , T I . Such boundary uncertainties stem most often from manufacturing uncertainties.

Appendix B. The nth-CASAM-N for n = 1, 2, 3

The model and boundary parameters α are considered to be uncertain quantities, with unknown true values. The nominal (or mean) parameter values α 0 are considered to be known, and these will differ from the true values by variations denoted as δ α δ α 1 , , δ α T P , where δ α i α i α i 0 . Since the forward state functions u x are related to the model and boundary parameters α through Equations (A1) and (A2), variations δ α in the model and boundary parameters will cause corresponding variations v 1 x δ u 1 x , , δ u T D x around the nominal solution u 0 x in the forward state functions. The variations δ α and v 1 x induce variations in the system’s response.
The first-order sensitivities of a model response R e , where e α , u E , with respect to variations h δ α , v 1 in the model parameters and state functions in a neighborhood around the nominal functions and parameter values e 0 α 0 , u 0 E , will exist if the first-order Gateaux (G) variation of the response will exist and will be linear in h δ α , v 1 , which will occur if and only if the following two conditions are satisfied by R e :
(i)
R e satisfies a weak Lipschitz condition at e 0 , namely,
R e 0 + ε h R e 0 k ε e 0 , k < ,
(ii)
R e satisfies the following condition for a scalar ε F , where the symbol F denotes the underlying field of scalars:
R e 0 + ε h 1 + ε h 2 R e 0 + ε h 1 R e 0 + ε h 2 + R e 0 = o ε ; h 1 , h 2 E .
Numerical methods (e.g., Newton’s method and variants thereof) for solving Equations (A1) and (A2) also require the existence of the first-order G-derivatives of original model equations. Therefore, the conditions provided in Equations (A10) and (A11) are henceforth considered to be satisfied by the model responses and also by the operators underlying the physical system modeled by Equations (A1) and (A2). When the first-order G-differential δ R e 0 ; h δ R u x ; α ; v 1 x ; δ α α 0 satisfies the conditions provided in Equations (A10) and (A11), it can be written as follows:
δ R u x ; α ; v 1 x ; δ α α 0 δ R u x ; α ; δ α d i r + δ R u x ; α ; v 1 x i n d .
In Equation (A12), the direct-effect term δ R u x ; α ; δ α d i r comprises only dependencies on δ α and is defined as follows:
δ R u x ; α ; δ α d i r R u ; α α α 0 δ α j 1 = 1 T P R 1 j 1 ; u x ; α d i r δ α j 1 ,
where R u ; α / α denotes the partial G-derivatives of R e with respect to α , evaluated at the nominal parameter values, and where the following definitions are used:
  α δ α i = 1 T P   α i δ α i ,
{ R ( 1 ) [ j 1 ; u ( x ) ; α ] } d i r { λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) S ( u ; α ; α ) α j 1 d x 1 d x T I } α 0 + j = 1 T I { λ 1 ( α ) ω 1 ( α ) λ j 1 ( α ) ω j 1 ( α ) λ j + 1 ( α ) ω j + 1 ( α ) λ T I ( α ) ω T I ( α ) S [ u ( x 1 , . , ω j ( α ) , . , x N x ) ; α ] ω j ( α ) α j 1 d x 1 d x T I } α 0 j = 1 T I { λ 1 ( α ) ω 1 ( α ) λ j 1 ( α ) ω j 1 ( α ) λ j + 1 ( α ) ω j + 1 ( α ) λ T I ( α ) ω T I ( α ) S [ u ( x 1 , . , λ j ( α ) , . , x N x ) ; α ] λ j ( α ) α j 1 d x 1 d x T I } α 0 .
The notation on the left side of Equation (A14) comprises an implied transposed vector (since it represents an inner product), but the respective dagger ( ), which indicates transposition, has been omitted in order to keep the notation as simple as possible. Daggers indicating transposition will also be omitted in other inner products whenever possible without causing ambiguities.
The direct-effect term can be computed once the nominal values e 0 = u 0 , α 0 are available. The notation α 0 will be used to indicate that the quantity enclosed within the bracket is to be evaluated at the respective nominal parameter and state function values.
On the other hand, the quantity δ R u x ; α ; v 1 x i n d in Equation (A12) comprises only variations in the state functions and is therefore called the indirect-effect term, with the following expression:
δ R u x ; α ; v 1 x i n d λ 1 α 0 ω 1 α 0 λ T I α 0 ω T I α 0 S u ; α ; x u α 0 v 1 x d x 1 d x T I ,
where
  u v 1 x i = 1 T D   u i x δ u i x .
The notation on the left side of Equation (A17) comprises an implied transposed vector (since it represents an inner product), but the respective dagger ( ), which indicates transposition, has been omitted to simplify the notation.
The indirect-effect term induces variations in the response through the variations in the state functions, which are, in turn, caused by the parameter variations through the equations underlying the model. Evidently, the indirect-effect term can be quantified only after having determined the variations v 1 x in terms of the variations δ α . The first-order relationship between the vectors v 1 x and δ α is determined by solving the equations obtained by applying the definition of the G-differential to Equations (A1) and (A2), which yields the following equations:
N 1 u ; α v 1 x α 0 = q V 1 u ; α ; δ α α 0 , x Ω x ,
b V 1 u ; α ; v 1 ; δ α α 0 = 0 , x Ω x α 0 .
In Equations (A18) and (A19), the superscript “(1)” indicates “first-level” and the various quantities that appear in these equations are defined as follows:
N 1 u ; α N u ; α u N 1 u 1 N 1 u T D N T D u 1 N T D u T D ;
q V 1 u ; α ; δ α Q α N u ; α α δ α j 1 = 1 T P s V 1 j 1 ; u ; α δ α j 1 ;
s V 1 j 1 ; u ; α Q α N u ; α α j 1 ;
b V 1 u ; α ; v 1 ; δ α α 0 B u ; α u α 0 v 1 + B u ; α C α α α 0 δ α .
The system comprising Equations (A81) and (A82) is called the 1st-Level Variational Sensitivity System (1st-LVSS). In order to determine the solutions of the 1st-LVSS that would correspond to every parameter variation δ α j 1 , j 1 = 1 , , T P , the 1st-LVSS would need to be solved T P times, with distinct right sides for each δ α j 1 , thus requiring T P large-scale computations. In other words, the actual form of the 1st-LVSS that would need to be solved in practice is as follows:
N 1 u ; α v 1 j 1 ; x α 0 = s V 1 j 1 ; u ; α α 0 , j 1 = 1 , , T P ; x Ω x ,
b V 1 u ; α ; v 1 j 1 ; x α 0 = 0 ; j 1 = 1 , , T P ; x Ω x α 0 .

Appendix B.1. The 1st-CASAM-N for Computing Exactly and Efficiently the 1st-Order Sensitivities R 1 j 1 ; u x ; a ( 1 ) x ; α R u x ; α / α j 1 , j 1 = 1 , , T P

The need for computing the vectors v 1 j 1 ; x , j 1 = 1 , , T P , is eliminated by expressing the indirect-effect term defined in Equation (A16) in terms of the solutions of the 1st-Level Adjoint Sensitivity System (1st-LASS), the construction of which requires the introduction of adjoint operators. This is accomplished by introducing a real Hilbert space, denoted as H 1 Ω x , endowed with an inner product of two vectors u ( a ) x H 1 and u ( b ) x H 1 , denoted as u ( a ) , u ( b ) 1 and defined as follows:
u ( a ) , u ( b ) 1 λ 1 α ω 1 α λ T I α ω T I α u ( a ) x u ( b ) x d x 1 d x T I α 0 ,
where u ( a ) x u ( b ) x i = 1 T D u i ( a ) x u i ( b ) x , and where the dagger ( ), which indicates transposition, has been omitted (to simplify the notation) in the representation of this scalar product.
Using the definition of the adjoint operator in H 1 Ω x , the left side of Equation (A81) is transformed as follows:
a ( 1 ) , N 1 u ; α v 1 1 α 0 = N 1 u ; α * a ( 1 ) , v 1 1 α 0 + P 1 u ; α ; a ( 1 ) ; v 1 Ω x α 0 ,
where P 1 u ; α ; a ( 1 ) ; v 1 Ω x denotes the associated bilinear concomitant evaluated on the space/time domain’s boundary Ω x α 0 and where N 1 u ; α * is the operator adjoint to N 1 u ; α . The symbol   indicates an adjoint operator. In certain situations, it might be computationally advantageous to include certain boundary components of P 1 u ; α ; a ( 1 ) ; v 1 Ω x in the components of N 1 u ; α * . The first term on the right side of Equation (A21) is required to represent the indirect-effect term defined in Equation (A16) by imposing the following relationship:
N 1 u ; α * a 1 x α 0 = S u ; α / u α 0 q A 1 u x ; α , x Ω x ,
The domain of A 1 u ; α is determined by selecting appropriate adjoint boundary and/or initial conditions, which will be denoted in operator form as:
b A 1 u ; a 1 ; α α 0 = 0 , x Ω x α 0 .
The above boundary conditions for N 1 u ; α * are usually inhomogeneous, i.e., b A 1 0 ; 0 ; α 0 , and are obtained by imposing the following requirements: (i) They must be independent of unknown values of v 1 x and δ α , and (ii) the substitution of the boundary and/or initial conditions represented by Equations (A25) and (A29) in the expression of P 1 u ; α ; a ( 1 ) ; v 1 Ω x α 0 must cause all terms containing unknown values of v 1 x to vanish. Constructing the adjoint initial and/or boundary conditions for N 1 u ; α * as described above and implementing them together with the variational boundary and initial conditions represented by Equation (A25) in Equation (A27) reduces the bilinear concomitant P 1 u ; α ; a ( 1 ) ; v 1 Ω x α 0 to a quantity denoted as P ^ 1 u ; α ; a ( 1 ) ; δ α Ω x α 0 , which will contain boundary terms involving only known values of δ α , α 0 , u 0 , and ψ ( 1 ) . Since P ^ 1 u ; α ; a ( 1 ) ; δ α Ω x α 0 is linear in δ α , it can be expressed in the following form: P ^ 1 u ; α ; a ( 1 ) ; δ α Ω x = j 1 = 1 T P P ^ 1 u ; α ; a ( 1 ) / α j 1 δ α j 1 .
The results obtained in Equations (A27) and (A28) are now replaced in Equation (A16) to obtain the following expression of the indirect-effect term as a function of a ( 1 ) x :
δ R u x ; α ; v 1 x i n d = a ( 1 ) , q V 1 u ; α ; δ α 1 α 0 P ^ 1 u ; α ; a ( 1 ) ; δ α Ω x α 0 ,
Replacing in Equation (A12) the result obtained in Equation (A30) together with the expression for the direct-effect term provided in Equation (A13) yields the following expression for the first G-differential of the response R u x ; α :
δ R u x ; α ; v 1 x ; δ α α 0 = δ R u x ; α ; δ α d i r + a ( 1 ) , q V 1 u ; α ; δ α 1 α 0 P ^ 1 u ; α ; a ( 1 ) ; δ α Ω x α 0 j 1 = 1 T P R 1 j 1 ; u x ; a ( 1 ) x ; α α 0 δ α j 1 ,
where for each j 1 = 1 , , T P , the quantity R 1 j 1 ; u x ; a 1 x ; α denotes the first-order sensitivities of the response R u x ; α with respect to the model parameters α j 1 and has the following expression:
R ( 1 ) [ j 1 ; u ( x ) ; a ( 1 ) ( x ) ; α ] = [ P ^ ( 1 ) ( u ; a ( 1 ) ; α ; ) ] Ω x α j 1 + { R ( 1 ) [ j 1 ; u ( x ) ; α ] } d i r + λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) a ( 1 ) ( x ) [ Q ( α ) N ( u ; α ) ] α j 1 d x 1 d x T I λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) S ( 1 ) [ j 1 ; u ( x ) ; a ( 1 ) ( x ) ; α ] d x 1 d x T I R [ u ( x ) ; α ] α j 1 ; j 1 = 1 , , T P .
As indicated by Equation (A32), each of the first-order sensitivities R 1 j 1 ; u x ; a 1 x ; α of the response R u x ; α with respect to the model parameters α j 1 (including boundary and initial conditions) can be computed inexpensively after having obtained the function a ( 1 ) x H 1 , using just quadrature formulas to evaluate the various inner products involving a ( 1 ) x H 1 . The function a ( 1 ) x H 1 is obtained by numerically solving Equations (A28) and (A29), which is the only large-scale computation needed to obtain all of the first-order sensitivities. Equations (A28) and (A29) are called the 1st-Level Adjoint Sensitivity System (1st-LASS), and its solution, a ( 1 ) x H 1 Ω x , is called the first-level adjoint function. It is very important to note that the 1st-LASS is independent of parameter variation δ α j 1 , j 1 = 1 , , T P , and therefore needs to be solved only once, regardless of the number of model parameters under consideration. Furthermore, since Equation (A28) is linear in a ( 1 ) x ψ 1 , i 1 2 x , solving it requires less computational effort than solving the original Equation (A1), which is nonlinear in u x . Finally, the function S 1 j 1 ; u x ; a 1 x ; α , which is defined and constructed by using the left side of Equation (A32), will contain Dirac-delta functions that are needed to represent the non-zero residual boundary terms contained in P ^ 1 u ; a ( 1 ) ; α ; as a volume integral over d x 1 d x T I . The function S 1 j 1 ; u x ; a 1 x ; α will be used to determine the second- and higher-order sensitivities of the response with respect to the model’s parameters.

Appendix B.2. The 2nd-CASAM-N for Computing Exactly and Efficiently the Second-Order Sensitivities, R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α 2 R u x ; α / α j 2 α j 1 , j 1 , j 2 = 1 , , T P

2nd-CASAM-N determines the second-order sensitivities by treating them as the first-order sensitivities of the first-order sensitivities. Thus, the first-order G-differential of a first-order sensitivity R 1 j 1 ; u x ; a 1 x ; α , j 1 = 1 , , T P is given by the following expression:
δ R 1 j 1 ; u x ; a 1 x ; α ; v 1 x ; δ a 1 x ; δ α α 0 = j 2 = 1 T P R 1 j 1 ; u ; a 1 ; α α j 2 α 0 δ α j 2 + δ R 1 j 1 ; u ; a 1 ; α ; v 1 x ; δ a 1 x i n d ,
where the indirect-effect term δ R 1 j 1 ; u ; a 1 ; α ; v 1 x ; δ a 1 x i n d comprises all dependencies on the vectors v 1 x and δ a 1 x of variations in the state functions u x and a 1 x , respectively, and is defined as follows:
δ R 1 j 1 ; u ; a 1 ; α ; v 1 ; δ a 1 i n d λ 1 α ω 1 α λ T I α ω T I α S 1 u v 1 x + S 1 a 1 δ a 1 x α 0 d x 1 d x T I .
The functions v 1 x and δ a 1 x are obtained by solving the following 2nd-Level Variational Sensitivity System (2nd-LVSS), which comprises the concatenation of the 1st-LVSS with the G-differentiated 1st-LASS:
V M 2 2 × 2 ; U 2 2 ; x ; α V 2 2 ; x α 0 = Q V 2 2 ; U 2 2 ; x ; α ; δ α α 0 , x Ω x ,
B V 2 2 ; U 2 2 ; x ; V 2 2 ; x ; α ; δ α α 0 = 0 2 , 0 2 0 , 0 , x Ω x α 0 .
The argument “2” that appears in the list of arguments of the vector U 2 2 ; x and the variational vector V 2 2 ; x in Equation (A35) indicate that each of these vectors is a two-block column vector, each block comprising a column vector of dimension T D , defined as follows:
U 2 2 ; x u 1 x a 1 x ; V 2 2 ; x δ U 2 2 ; x v 2 1 ; x v 2 2 ; x v 1 x δ a 1 x .
To distinguish block vectors from block matrices, two bold capital letters have been used (and will henceforth be used) to denote block matrices, as in the case of the second-level variational matrix V M 2 2 × 2 ; u 2 x ; α . The second level is indicated by the superscript “(2)”. The argument 2 × 2 , which appears in the list of arguments of V M 2 2 × 2 ; u 2 x ; α , indicates that this matrix is a 2 × 2 -dimensional block matrix comprising four matrices, each of dimensions T D × T D , with the following structure:
V M 2 2 × 2 ; U 2 2 ; x ; α N 1 0 V 21 2 V 22 2 .
The other quantities that appear in Equations (A35) and (A36) are two-block vectors with the same structure as V 2 2 ; x and are defined as follows:
Q V 2 2 ; U 2 2 ; x ; α ; δ α q V 2 1 ; U 2 2 ; x ; α ; δ α q V 2 2 ; U 2 2 ; x ; α ; δ α q V 1 u ; α ; δ α q 2 2 u ; a 1 ; α ; δ α ;
q 2 2 u ; α ; a 1 ; δ α q A 1 u x ; α α δ α N 1 u ; α * a 1 x α δ α ;
q V 2 i ; U 2 2 ; x ; α ; δ α j 2 = 1 T P s V 2 i ; j 2 ; U 2 2 ; x ; α δ α j 2 , i = 1 , 2 ;
s V 2 1 ; j 2 ; U 2 2 ; x ; α Q α N u ; α α j 2 ;
s V 2 2 ; j 2 ; U 2 2 ; x ; α q A 1 u x ; α α j 2 N 1 u ; α * a 1 x α j 2 ;
B V 2 2 ; U 2 2 ; x ; V 2 2 ; x ; α ; δ α b V 2 1 ; U 2 2 ; x ; V 2 2 ; x ; α ; δ α b V 2 2 ; U 2 2 ; x ; V 2 2 ; x ; α ; δ α b V 1 u 1 ; α ; δ u 1 ; δ α δ b A 1 U 2 2 ; x ; V 2 2 ; x ; α ; δ α .
V 21 2 u ; a 1 ; α N 1 u ; α * a 1 u q A 1 u x ; α u ;
V 22 2 u ; α N 1 u ; α * ;
δ b A 1 u ; a 1 ; α b A 1 u v 1 x + b A 1 a 1 δ a 1 x + b A 1 α δ α .
The need to solve the 2nd-LVSS is circumvented by deriving an alternative expression for the indirect-effect term δ R 1 j 1 ; u ; a 1 ; α ; V 2 2 ; x i n d defined in Equation (A34), in which the function V 2 2 ; x is replaced by a second-level adjoint function that is independent of variations in the model parameter and state functions and is the solution of a 2nd-Level Adjoint Sensitivity System (2nd-LASS) that is constructed by using the same principles as employed for deriving the 1st-LASS.
The 2nd-LASS is constructed in a Hilbert space, denoted as H 2 Ω x , which comprises as elements block vectors of the same form as V 2 2 ; x , and is endowed with the following inner product of two vectors Ψ ( 2 ) 2 ; x ψ ( 2 ) 1 ; x , ψ ( 2 ) 2 ; x H 2 Ω x and Φ ( 2 ) x φ ( 2 ) 1 ; x , φ ( 2 ) 2 ; x H 2 Ω x :
Ψ ( 2 ) 2 ; x , Φ ( 2 ) x 2 i = 1 2 ψ ( 2 ) i ; x , φ ( 2 ) i ; x 1 .  
The inner product defined in Equation (A48) is used to construct the following 2nd-Level Adjoint Sensitivity System (2nd-LASS) for the second-level adjoint function A ( 2 ) 2 ; j 1 ; x a ( 2 ) 1 ; j 1 ; x , a ( 2 ) 2 ; j 1 ; x H 2 Ω x for each j 1 = 1 , , T P :
A M 2 2 × 2 ; U 2 2 ; x ; α A ( 2 ) 2 ; j 1 ; x α 0 = Q A 2 2 ; j 1 ; U 2 2 ; x ; α α 0 , j 1 = 1 , , T P ; x Ω x ,
subject to boundary conditions represented as follows:
B A 2 2 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α α 0 = 0 2 ; j 1 = 1 , , T P ; x Ω x α 0 .
where
Q A 2 2 ; j 1 ; U 2 2 ; x ; α q A 2 1 ; j 1 ; U 2 ; α q A 2 2 ; j 1 ; U 2 ; α S 1 j 1 ; u x ; a 1 x ; α ; v 1 x / u S 1 j 1 ; u x ; a 1 x ; α ; v 1 x / a 1 , j 1 = 1 , , T P .
A M 2 2 × 2 ; U 2 2 ; x ; α V M 2 2 × 2 ; U 2 2 ; x ; α * = N 1 u ; α * V 21 2 * 0 V 22 2 * ;
The matrix A M 2 2 × 2 ; u 2 x ; α comprises 2 × 2 block matrices, each of dimensions T D 2 , thus comprising a total of 2 × 2 T D 2  components (or elements), and is obtained from the following relation:
A ( 2 ) 2 ; x , V M 2 V 2 2 ; x 2 α 0 = P 2 U 2 ; A ( 2 ) ; V 2 ; α Ω x α 0 + V 2 2 ; x , A M 2 2 × 2 ; U 2 2 ; x ; α A ( 2 ) 2 ; x 2 α 0 ,
where the quantity P 2 U 2 ; A ( 2 ) ; V 2 ; α Ω x α 0 denotes the corresponding bilinear concomitant on the domain’s boundary, evaluated at the nominal values for the parameters and respective state functions. The second-level adjoint boundary/initial conditions represented by Equation (A50) are determined by requiring that (a) they be independent of unknown values of V 2 2 ; x , and (b) the substitution of the boundary and/or initial conditions represented by Equations (A50) and (A36) in the expression of P 2 U 2 ; A ( 2 ) ; V 2 ; α Ω x α 0 cause all terms containing unknown values of V 2 2 ; x to vanish. Using in Equation (A34) the relations defining 2nd-LASS, the 2nd-LVSS and the relation provided in Equation (A53) yields the following alternative expression for the indirect-effect term in terms of the second-level adjoint sensitivity function A ( 2 ) 2 ; j 1 ; x a ( 2 ) 1 ; j 1 ; x , a ( 2 ) 2 ; j 1 ; x :
δ R 1 j 1 ; u ; a 1 ; α ; V 2 2 ; x i n d = P ^ 2 U 2 ; A ( 2 ) ; α ; δ α Ω x α 0 + A ( 2 ) 2 ; j 1 ; x , Q V 2 2 ; U 2 2 ; x ; α ; δ α 2 α 0 ,
where P ^ 2 U 2 ; A ( 2 ) ; α ; δ α Ω x α 0  denotes residual boundary terms that may not have vanished after having used the boundary and/or initial conditions represented by Equations (A36) and (A50). Replacing the expression obtained in Equation (A54) in Equation (A33) yields the following expression:
δ R 1 j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α ; δ α α 0 = = j 2 = 1 T P R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α α 0 δ α j 2 , j 1 = 1 , , T P ,
where the quantity R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α denotes the second-order sensitivity of the generic scalar-valued response R u x ; α with respect to the parameters α j 1 and α j 2 computed at the nominal values of the parameters and respective state functions, and has the following expression for j 1 , j 2 = 1 , , T P :
R ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ] R ( 1 ) [ j 1 ; u ( x ) ; a ( 1 ) ( x ) ; α ] α j 2 { P ^ ( 2 ) [ U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ] } Ω x α j 2 + i = 1 2 a ( 2 ) ( i ; j 1 ; x ) , s V ( 2 ) [ i ; j 2 ; U ( 2 ) ( 2 ; x ) ; α ] 1 λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) S ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ] d x 1 d x T I 2 R [ u ( x ) ; α ] α j 2 α j 1 .
If the 2nd-LASS is solved T P times, the second-order mixed sensitivities R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α 2 R / α j 2 α j 1 will be computed twice, in two different ways, in terms of two distinct second-level adjoint functions. Consequently, the symmetry property 2 R u x ; α / α j 2 α j 1 = 2 R u x ; α / α j 1 α j 2 enjoyed by the second-order sensitivities provides an intrinsic (numerical) verification that the components of the second-level adjoint function A ( 2 ) 2 ; j 1 ; x , as well as the first-level adjoint function a ( 1 ) x , are computed accurately.

Appendix B.3. The 3rd-CASAM-N for Computing Exactly and Efficiently the 3rd-Order Sensitivities, R 3 j 3 ; j 2 ; j 1 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α 3 R u x ; α / α j 3 α j 2 α j 1 , j 1 , j 2 , j 3 = 1 , , T P

The third-order sensitivities are computed by considering them to be the first-order sensitivities of a second-order sensitivity. Thus, each of the second-order sensitivities R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α 2 R / α j 2 α j 1 will be considered to be a model response that is assumed to satisfy the conditions stated in Equations (A10) and (A11) for each j 1 , j 2 = 1 , , T P , so that the first-order total G-differential of R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α will exist and will be linear in the variations V 2 2 ; x and δ A ( 2 ) 2 ; j 1 ; x in a neighborhood around the nominal values of the parameters and the respective state functions. By definition, the first-order total G-differential of R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α , which will be denoted as δ R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α ; V 2 2 ; x ; δ A ( 2 ) 2 ; j 1 ; x ; δ α α 0 , is given by the following expression:
{ δ R ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ; V ( 2 ) ( 2 ; x ) ; δ A ( 2 ) ( 2 ; j 1 ; x ) ; δ α ] } α 0 { R ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ] α δ α } α 0 + { δ R ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ; V ( 2 ) ( 2 ; x ) ; δ A ( 2 ) ( 2 ; j 1 ; x ) ] } i n d ,
where:
{ δ R ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ( 2 ; x ) ; A ( 2 ) ( 2 ; j 1 ; x ) ; α ; V ( 2 ) ( 2 ; x ) ; δ A ( 2 ) ( 2 ; j 1 ; x ) ] } i n d λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) { S ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ; A ( 2 ) ; α ] U ( 2 ) ( 2 ; x ) V ( 2 ) ( 2 ; x ) } α 0 d x 1 d x T I + λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) { S ( 2 ) [ j 2 ; j 1 ; U ( 2 ) ; A ( 2 ) ; α ] A ( 2 ) ( 2 ; j 1 ; x ) δ A ( 2 ) ( 2 ; j 1 ; x ) } α 0 d x 1 d x T I .
The indirect-effect term δ R 2 j 2 ; j 1 ; U 2 2 ; x ; A ( 2 ) 2 ; j 1 ; x ; α ; δ α d i r can be computed after having determined the vectors V 2 2 ; x and δ A ( 2 ) 2 ; j 1 ; x , which are the solutions of the following 3rd-Level Variational Sensitivity System (3rd-LVSS):
V M 3 4 × 4 ; U 3 4 ; j 1 ; x ; α V 3 4 ; j 1 ; x α 0 = Q V 3 4 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α α 0 , x Ω x ,
B V 3 4 ; U 3 4 ; j 1 ; x ; V 3 4 ; j 1 ; x ; α ; δ α α 0 = 0 4 ; 0 4 0 , 0 , 0 , 0 , x Ω x α 0
where
V M 3 4 × 4 ; U 3 ; α V M 2 2 × 2 0 2 × 2 V M 21 3 2 × 2 V M 22 3 2 × 2 ;
U 3 4 ; j 1 ; x U 2 2 ; x A ( 2 ) 2 ; j 1 ; x ; V 3 4 ; j 1 ; x δ U 3 4 ; j 1 ; x = V 2 2 ; x δ A ( 2 ) 2 ; j 1 ; x ;
V M 21 3 2 × 2 ; x A M 2 2 × 2 ; U 2 2 ; x ; α A ( 2 ) 2 ; j 1 ; x U 2 2 ; x Q A 2 2 ; j 1 ; u 2 x ; α U 2 2 ; x ;
V M 22 3 2 × 2 ; x A M 2 2 × 2 ; U 2 2 ; x ; α ; 0 2 × 2 0 0 0 0 ;
Q V 3 4 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α Q V 2 2 ; U 2 2 ; x ; α ; δ α Q 2 3 2 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α q V 3 1 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α , , q V 3 4 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α ;
q V 3 i ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α j 3 = 1 T P s V 3 i ; j 3 ; j 1 ; U 3 4 ; j 1 ; x ; α δ α j 3 ; i = 1 , ...4 ;
Q 2 3 2 ; j 1 ; U 3 4 ; j 1 ; x ; α ; δ α Q A 2 2 ; j 1 ; u 2 x ; α α α A M 2 2 × 2 ; U 2 2 ; x ; α A ( 2 ) 2 ; j 1 ; x α α ;
B V 3 4 ; U 3 4 ; j 1 ; x ; V 3 4 ; j 1 ; x ; α ; δ α B V 2 2 ; U 2 2 ; x ; V 2 2 ; x ; α ; δ α δ B A 2 2 ; U 3 4 ; j 1 ; x ; V 3 4 ; j 1 ; x ; α ; δ α
Solving the 3rd-LVSS would require T P 3 large-scale computations, which is unrealistic for large-scale systems comprising many parameters. The 3rd-CASAM-N circumvents the need to solve the 3rd-LVSS by deriving an alternative expression for the indirect-effect term defined in Equation (A58), in which the function V 3 4 ; j 1 ; x is replaced by a third-level adjoint function that is independent of parameter variations. This third-level adjoint function is the solution of a 3rd-Level Adjoint Sensitivity System (3rd-LASS), which is constructed by applying the same principles as those used for constructing the 1st-LASS and the 2nd-LASS. The Hilbert space appropriate for constructing the 3rd-LASS, denoted as H 3 Ω x , comprises as elements block vectors of the same form as V 3 4 ; j 1 ; x . Thus, a generic block vector in H 3 Ω x , denoted as Ψ ( 3 ) 4 ; x ψ ( 3 ) 1 ; x , ψ ( 3 ) 2 ; x , ψ ( 3 ) 3 ; x , ψ ( 3 ) 4 ; x H 3 Ω x , comprises four T D -dimensional vector components of the form ψ 3 i ; x ψ 1 3 i ; x , , ψ T D 3 i ; x H 1 Ω x , i = 1 , 2 , 3 , 4 , where each of these four components is a T D -dimensional column vector. The inner product of two vectors Ψ ( 3 ) 4 ; x H 3 Ω x and Φ ( 3 ) 4 ; x H 3 Ω x in the Hilbert space H 3 Ω x will be denoted as Ψ ( 3 ) 4 ; x , Φ ( 3 ) 4 ; x 3  and defined as follows:
Ψ ( 3 ) 4 ; x , Φ ( 3 ) 4 ; x 3 i = 1 4 ψ ( 3 ) i ; x , φ ( 3 ) i ; x 1 .
The steps for constructing the 3rd-LASS are conceptually similar to those underlying the construction of the 1st-LASS and 2nd-LASS. The final expressions for the third-order sensitivity results are as follows:
δ R 2 j 2 ; j 1 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α ; δ α α 0 = R 2 j 2 ; j 1 ; U 3 ; α α δ α α 0 P ^ 3 U 3 ; A ( 3 ) ; δ α Ω x α 0 + A ( 3 ) 4 ; j 2 ; j 1 ; x , Q V 3 4 ; j 1 ; U 3 ; α ; δ α 3 α 0 ,
where P ^ 3 U 3 ; A ( 3 ) ; δ α Ω x α 0 denotes residual boundary terms that may have not vanished automatically and where the third-level adjoint function A ( 3 ) 4 ; x a ( 3 ) 1 ; x , a ( 3 ) 2 ; x , a ( 3 ) 3 ; x , a ( 3 ) 4 ; x H 3 Ω x  is the solution of the following 3rd-LASS:
A M 3 4 × 4 ; U 3 4 ; j 1 ; x ; α A ( 3 ) 4 ; j 2 ; j 1 ; x α 0 = Q A 3 4 ; j 2 ; j 1 ; U 3 4 ; j 1 ; x ; α α 0 , j 1 = 1 , , T P ; j 2 = 1 , , j 1 ,
where
Q A 3 4 ; j 2 ; j 1 ; U 3 4 ; j 1 ; x ; α q A 3 1 ; j 2 ; j 1 ; U 3 ; α , , q A 3 4 ; j 2 ; j 1 ; U 3 ; α ;
q A 3 1 ; j 2 ; j 1 ; U 3 ; α S 2 j 2 ; j 1 ; u 2 ; a 2 ; α / u 1 ;
q A 3 2 ; j 2 ; j 1 ; U 3 ; α S 2 j 2 ; j 1 ; u 2 ; a 2 ; α / a 1 ;
q A 3 3 ; j 2 ; j 1 ; U 3 ; α S 2 j 2 ; j 1 ; u 2 ; a 2 ; α / a ( 2 ) 1 ; j 1 ; x ;
q A 3 4 ; j 2 ; j 1 ; U 3 ; α S 2 j 2 ; j 1 ; u 2 ; a 2 ; α / a ( 2 ) 2 ; j 1 ; x .
A M 3 4 × 4 ; U 3 4 ; j 1 ; x ; α V M 3 4 × 4 ; U 3 ; α * = V M 2 2 × 2 * V M 21 3 2 × 2 * 0 2 × 2 V M 22 3 2 × 2 * ,
The boundary conditions to be satisfied by each of the third-level adjoint functions A ( 3 ) 4 ; j 2 ; j 1 ; x a ( 3 ) 1 ; j 2 ; j 1 ; x , a ( 3 ) 2 ; j 2 ; j 1 ; x , a ( 3 ) 3 ; j 2 ; j 1 ; x , a ( 3 ) 4 ; j 2 ; j 1 ; x  can be represented in operator form as follows:
B A 3 4 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α α 0 = 0 4 ; f o r j 1 = 1 , , T P ; j 2 = 1 , , j 1 ; x Ω x α 0 .
In component form, the total differential expressed by Equation (A70) has the following expression:
δ R 2 j 2 ; j 1 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α ; δ α α 0 = j 3 = 1 T P R 3 j 3 ; j 2 ; j 1 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α α 0 δ α j 3 , j 1 ; j 2 = 1 , , T P ,
where the quantity R 3 j 3 ; j 2 ; j 1 ; U 3 4 ; j 1 ; x ; A ( 3 ) 4 ; j 2 ; j 1 ; x ; α denotes the third-order sensitivity of the generic scalar-valued response R u x ; α with respect to any three model parameters α j 1 , α j 2 , α j 3 , and has the following expression for j 1 , j 2 , j 3 = 1 , , T P :
R ( 3 ) [ j 3 ; j 2 ; j 1 ; U ( 3 ) ( 4 ; j 1 ; x ) ; A ( 3 ) ( 4 ; j 2 ; j 1 ; x ) ; α ] R ( 2 ) [ j 2 ; j 1 ; U ( 3 ) ( 4 ; j 1 ; x ) ; α ] α j 3 [ P ^ ( 3 ) ( U ( 3 ) ; A ( 3 ) ; δ α ) ] Ω x α j 3 + i = 1 4 a ( 3 ) ( i ; j 2 ; j 1 ; x ) , s V ( 3 ) [ i ; j 3 ; j 1 ; U ( 3 ) ( 4 ; j 1 ; x ) ; α ] 1 λ 1 ( α ) ω 1 ( α ) λ T I ( α ) ω T I ( α ) S ( 3 ) [ j 3 ; j 2 ; j 1 ; U ( 3 ) ( 4 ; j 1 ; x ) ; A ( 3 ) ( 4 ; j 2 ; j 1 ; x ) ; α ] d x 1 d x T I 3 R [ u ( x ) ; α ] α j 3 α j 2 α j 1 .

Appendix B.4. The Mathematical Framework of the 1st-CASAM-N in Block Matrix/Vector Notation

It is apparent from the formulations of the 2nd-CASAM-N and 3rd-CASAM-N that the general mathematical framework of the nth-CASAM-N will involve block matrices and block vectors. The 1st-CASAM-N, however, only involves T D -dimensional vectors and matrices but does not involve block matrices or block vectors. In order to facilitate the comparison of the formulas obtained when particularizing the general mathematical framework of the nth-CASAM-N for the particular case n = 1 , it is convenient to recast the formulas obtained for the 1st-CASAM-N in block-vector and block-matrix form of the types used in the formulation of the 2nd- and 3rd-CASAM-N. Thus, Equations (A18) and (A19) take on the following forms in block-vector/-matrix notation:
V M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α V 1 2 0 ; x α 0 = Q V 1 2 0 ; U 1 2 0 ; x ; α ; δ α α 0 , x Ω x ,
B V 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α α 0 = 0 , x Ω x α 0 .
In Equations (A81) and (A82), the superscript “(1)” indicates “first-level,” 2 0 1  (evidently), and the various quantities that appear in these equations are defined as follows:
U 1 2 0 ; x u x ; V 1 2 0 ; x v 1 x ;
V M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α N 1 u ; α N u ; α u ;
Q V 1 2 0 ; U 1 2 0 ; x ; α ; δ α q V 1 u ; α ; δ α Q α N u ; α α δ α ; ;
B V 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α b V 1 u ; α ; v 1 ; δ α .
In block-vector/-matrix form, the 1st-LASS has the following expression:
A M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α A 1 2 0 ; x α 0 = Q A 1 2 0 ; U 1 2 0 ; x ; α ; δ α α 0 , x Ω x ,
B A 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α α 0 = 0 , x Ω x α 0 .
where
A 1 2 0 ; x a 1 x ; A M 1 2 0 × 2 0 ; U 1 2 0 ; x ; α N 1 u ; α * ;
Q A 1 2 0 ; U 1 2 0 ; x ; α ; δ α q A 1 u x ; α S u ; α / u α 0 ;
B V 1 2 0 ; U 1 2 0 ; x ; V 1 2 0 ; x ; α ; δ α b A 1 u ; a 1 ; α .

References

  1. Bellman, R.E. Dynamic Programming; Rand Corporation and Princeton University Press: Princeton, NJ, USA, 1957; ISBN 978-0-691-07951-6. [Google Scholar]
  2. Wigner, E.P. Effect of Small Perturbations on Pile Period, Chicago Report CP-G-3048; Chicago. Univ. Metallurgical Lab.: Chicago, IL, USA, 1945. [Google Scholar]
  3. Cacuci, D.G. Sensitivity Theory for Nonlinear Systems: I. Nonlinear Functional Analysis Approach. J. Math. Phys. 1981, 22, 2794–2802. [Google Scholar] [CrossRef]
  4. Cacuci, D.G. Sensitivity Theory for Nonlinear Systems: II. Extensions to Additional Classes of Responses. J. Math. Phys. 1981, 22, 2803–2812. [Google Scholar] [CrossRef]
  5. Práger, T.; Kelemen, F.D. Adjoint methods and their application in earth sciences. In Advanced Numerical Methods for Complex Environmental Models: Needs and Availability; Faragó, I., Havasi, Á., Zlatev, Z., Eds.; Bentham Science Publishers: Oak Park, IL, USA, 2014; Chapter 4A; pp. 203–275. [Google Scholar]
  6. Luo, Z.; Wang, X.; Liu, D. Prediction on the static response of structures with large-scale uncertain-but-bounded parameters based on the adjoint sensitivity analysis. Struct. Multidiscip. Optim. 2020, 61, 123–139. [Google Scholar] [CrossRef]
  7. Cacuci, D.G. Second-Order Adjoint Sensitivity Analysis Methodology (2nd-ASAM) for Computing Exactly and Efficiently First- and Second-Order Sensitivities in Large-Scale Linear Systems: I. Computational Methodology. J. Comp. Phys. 2015, 284, 687–699. [Google Scholar] [CrossRef] [Green Version]
  8. Cacuci, D.G. Second-Order Adjoint Sensitivity Analysis Methodology (2nd-ASAM) for Large-Scale Nonlinear Systems: I. Theory. Nucl. Sci. Eng. 2016, 184, 16–30. [Google Scholar] [CrossRef]
  9. Cacuci, D.G.; Fang, R.; Favorite, J.A. Comprehensive Second-Order Adjoint Sensitivity Analysis Methodology (2nd-ASAM) Applied to a Subcritical Experimental Reactor Physics Benchmark: VI. Overall Impact of 1st- and 2nd-Order Sensitivities. Energies 2020, 13, 1674. [Google Scholar] [CrossRef] [Green Version]
  10. Valentine, T.E. Polyethylene-reflected plutonium metal sphere subcritical noise measurements, SUB-PU-METMIXED-001. In International Handbook of Evaluated Criticality Safety Benchmark Experiments, NEA/NSC/DOC(95)03/I-IX; Organization for Economic Co-Operation and Development (OECD), Nuclear Energy Agency (NEA): Paris, France, 2006. [Google Scholar]
  11. Fang, R.; Cacuci, D.G. Fourth-Order Adjoint Sensitivity and Uncertainty Analysis of an OECD/NEA Reactor Physics Benchmark: I. Computed Sensitivities. J. Nucl. Eng. 2021, 2, 281–308. [Google Scholar] [CrossRef]
  12. Cacuci, D.G. The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Response-Coupled Forward/Adjoint Linear Systems (nth-CASAM-L): I. Mathematical Framework. Energies 2021, 14, 8314. [Google Scholar] [CrossRef]
  13. Levine, H.; Schwinger, J. On the theory of diffraction by an aperture in an infinite plane screen. Phys. Rev. 1949, 75, 1423. [Google Scholar] [CrossRef]
  14. Roussopolos, P. Methodes variationeles en theories des collisions. Comptes Rendus Acad. Sci. 1953, 236, 1858. [Google Scholar]
  15. Lewins, J. IMPORTANCE: The Adjoint Function; Pergamon Press Ltd.: Oxford, UK, 1965. [Google Scholar]
  16. Stacey, W.M. Variational Methods in Nuclear Reactor Physics; Academic Press, Inc.: New York, NY, USA, 1974. [Google Scholar]
  17. Cacuci, D.G. The Fourth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (4th-CASAM-N): I. Mathematical Framework. J. Nucl. Eng. 2022, 3, 37–71. [Google Scholar] [CrossRef]
  18. Cacuci, D.G. The Fifth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (5th-CASAM-N): I. Mathematical Framework. AJCM 2022, 12, 44–78. [Google Scholar] [CrossRef]
  19. Cacuci, D.G. On the Need to Determine Accurately the Impact of Higher-Order Sensitivities on Model Sensitivity Analysis, Uncertainty Quantification and Best-Estimate Predictions. Energies 2021, 14, 6318. [Google Scholar] [CrossRef]
  20. Cacuci, D.G. BERRU Predictive Modeling: Best Estimate Results with Reduced Uncertainties; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 2019. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cacuci, D.G. The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N): Mathematical Framework. J. Nucl. Eng. 2022, 3, 163-190. https://doi.org/10.3390/jne3030010

AMA Style

Cacuci DG. The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N): Mathematical Framework. Journal of Nuclear Engineering. 2022; 3(3):163-190. https://doi.org/10.3390/jne3030010

Chicago/Turabian Style

Cacuci, Dan Gabriel. 2022. "The nth-Order Comprehensive Adjoint Sensitivity Analysis Methodology for Nonlinear Systems (nth-CASAM-N): Mathematical Framework" Journal of Nuclear Engineering 3, no. 3: 163-190. https://doi.org/10.3390/jne3030010

Article Metrics

Back to TopTop