scispace - formally typeset
Open AccessJournal ArticleDOI

Interior Methods for Nonlinear Optimization

Reads0
Chats0
TLDR
A condensed, selective look at classical material and recent research about interior methods for nonlinearly constrained optimization shows how their influence has transformed both the theory and practice of constrained optimization.
Abstract
Interior methods are an omnipresent, conspicuous feature of the constrained optimization landscape today, but it was not always so. Primarily in the form of barrier methods, interior-point techniques were popular during the 1960s for solving nonlinearly constrained problems. However, their use for linear programming was not even contemplated because of the total dominance of the simplex method. Vague but continuing anxiety about barrier methods eventually led to their abandonment in favor of newly emerging, apparently more efficient alternatives such as augmented Lagrangian and sequential quadratic programming methods. By the early 1980s, barrier methods were almost without exception regarded as a closed chapter in the history of optimization. This picture changed dramatically with Karmarkar's widely publicized announcement in 1984 of a fast polynomial-time interior method for linear programming; in 1985, a formal connection was established between his method and classical barrier methods. Since then, interior methods have advanced so far, so fast, that their influence has transformed both the theory and practice of constrained optimization. This article provides a condensed, selective look at classical material and recent research about interior methods for nonlinearly constrained optimization.

read more

Content maybe subject to copyright    Report

SIAM REVIEW
c
2002 Society for Industrial and Applied Mathematics
Vol. 44, No. 4, pp. 525–597
Interior Methods for Nonlinear
Optimization
Anders Forsgren
Philip E. Gill
Margaret H. Wright
§
Abstract. Interior methods are an omnipresent, conspicuous feature of the constrained optimiza-
tion landscape today, but it was not always so. Primarily in the form of barrier methods,
interior-point techniques were popular during the 1960s for solving nonlinearly constrained
problems. However, their use for linear programming was not even contemplated because
of the total dominance of the simplex method. Vague but continuing anxiety about barrier
methods eventually led to their abandonment in favor of newly emerging, apparently more
efficient alternatives such as augmented Lagrangian and sequential quadratic programming
methods. By the early 1980s, barrier methods were almost without exception regarded as
a closed chapter in the history of optimization.
This picture changed dramatically with Karmarkar’s widely publicized announcement
in 1984 of a fast polynomial-time interior method for linear programming; in 1985, a formal
connection was established between his method and classical barrier methods. Since then,
interior methods have advanced so far, so fast, that their influence has transformed both
the theory and practice of constrained optimization. This article provides a condensed, se-
lective look at classical material and recent research about interior methods for nonlinearly
constrained optimization.
Key words. nonlinear programming, constrained minimization, nonlinear constraints, primal-dual
methods, interior methods, penalty methods, barrier methods
AMS subject classifications. 49J20, 49J15, 49M37, 49D37, 65F05, 65K05, 90C30
PII. S0036144502414942
1. Introduction. It is a truth universally acknowledged that the field of continu-
ous optimization has undergone a dramatic change since 1984. This change, sometimes
described as the “interior-point revolution,” has featured a continual blending of old
and new, with effects far beyond optimization. An especially appealing aspect of the
interior-point revolution is its spirit of unification, which has brought together areas
of optimization that for many years were treated as firmly disjoint. Prior to 1984,
linear and nonlinear programming, one a subset of the other, had evolved for the
Received by the editors September 22, 2002; accepted for publication (in revised form) October 1,
2002; published electronically October 30, 2002.
http://www.siam.org/journals/sirev/44-4/41494.html
Optimization and Systems Theory, Department of Mathematics, Royal Institute of Technol-
ogy, SE-100 44 Stockholm, Sweden (anders.forsgren@math.kth.se). The research of this author was
supported by the Swedish Research Council (VR).
Department of Mathematics, University of California, San Diego, La Jolla, CA 92093-0112
(pgill@ucsd.edu). The research of this author was supported by National Science Foundation grants
DMS-0208449 and ACI-0082100.
§
Department of Computer Science, Courant Institute, New York University, New York, NY 10012
(mhw@cs.nyu.edu).
525

526 ANDERS FORSGREN, PHILIP E. GILL, AND MARGARET H. WRIGHT
most part along unconnected paths, without even a common terminology. (The use
of “programming” to mean “optimization” serves as a persistent reminder of these
differences.) Today this separation seems, as it indeed was, artificial, yet it was a fully
accepted part of the culture of optimization not so many years ago.
1.1. Roots in Linear and Nonlinear Programming. Although the focus of this
article is on nonlinearly constrained problems, understanding the context of the
interior-point revolution requires a short digression on linear programming (mini-
mization of a linear function subject to linear constraints). A fundamental property
of well-behaved n-variable linear programs with m inequality constraints is that a ver-
tex minimizer must exist, i.e., a point where n constraints with linearly independent
gradients hold with equality. (See, e.g., [20, 92] for details about linear program-
ming.) The simplex method, invented by Dantzig in 1947, is an iterative procedure
that solves linear programs by exploiting this property. A simplex iteration moves
from vertex to vertex, changing (one at a time) the set of constraints that hold ex-
actly, decreasing the objective as it goes, until an optimal vertex is found. From the
very start, the simplex method dominated the field of linear programming. Although
“nonsimplex” strategies for linear programming were suggested and tried from time
to time, they could not consistently match the simplex method in overall speed and
reliability. Furthermore, a simplex-centric world view had the effect that even “new”
techniques mimicked the motivation of the simplex method by always staying on a
subset of exactly satisfied constraints.
The preeminence of the simplex method was challenged not because of failures
in practice—the simplex method was, and is, used routinely to solve enormous linear
programs—but by worries about its computational complexity. One can argue that
the simplex method and its progeny are inherently combinatorial, in that their per-
formance seems to be bound in the worst case to the maximum number of ways in
which n out of m constraints can hold with equality. In fact, with standard pivoting
rules specifying the constraint to be dropped and added at each iteration, the simplex
method can visit every vertex of the feasible region [64]; thus its worst-case complex-
ity is exponential in the problem dimension. As a result, there was great interest in
finding a polynomial-time linear programming algorithm.
1
The first success in this direction was achieved in 1979 by Khachian, whose el-
lipsoid method was derived from approaches proposed originally for nonlinear opti-
mization. (See [92] for details about Khachian’s method.) Despite its polynomial
complexity bound, however, the ellipsoid method performed poorly in practice com-
pared to the simplex method, and the search continued for a polynomial-time linear
programming method that was genuinely fast in running time.
The start of the interior-point revolution was Karmarkar’s announcement [63]
in 1984 of a polynomial-time linear programming method that was 50 times faster
than the simplex method. Amid the frenzy of interest in Karmarkar’s method, it
was shown in 1985 [51] that there was a formal equivalence between Karmarkar’s
method and the classical logarithmic barrier method (see sections 1.2 and 3) applied
to linear programming, and long-discarded barrier methods were soon rejuvenated
as polynomial-time algorithms for linear programming. Furthermore, barrier meth-
ods (unlike the simplex method) could be applied not only to linear programming
1
Assuming various distributions of random inputs, [11, 94] showed that the simplex method
converges in expected polynomial time. The recent development of “smoothed” complexity analysis
[95] has led to new insights about the average behavior of the simplex method.

INTERIOR METHODS 527
but also to other optimization problems, such as quadratic programming, linear and
nonlinear complementarity, and nonlinear programming. Although the ties between
Karmarkar’s method and barrier methods were controversial for a few years, these
disagreements have (mostly) faded away. Pedagogical and philosophical issues remain
about the best way to motivate interior-point methods—perturbing optimality condi-
tions? minimizing a barrier function?—and the multiplicity of viewpoints continues
to create new insights and new algorithms.
The interior-point revolution has led to a fundamental shift in thinking about
continuous optimization. Linear and nonlinear programming are seen as related parts
of a unified whole, and no one would seriously claim today that linear programming
is completely different from nonlinear optimization. (Of course, methods for solving
linear programs and nonlinear problems vary significantly in detail.)
As we shall see, the signature of interior methods is the existence of continu-
ously parameterized families of approximate solutions that asymptotically converge
to the exact solution. These paths trace smooth trajectories with algebraic and geo-
metric properties (such as being “centered” in a precisely defined sense) that can be
analyzed and exploited algorithmically. Many interior methods are characterized as
“path-following” to signal their dependence on properties of these paths, which pro-
vide the foundation for all complexity analyses of interior-point algorithms for linear,
quadratic, and convex programming.
The monumental work [79] of Nesterov and Nemirovskii proposed new families
of barrier methods and extended polynomial-time complexity results to new convex
optimization problems. Semidefinite programming—minimization of a convex func-
tion in the space of symmetric matrices subject to semidefiniteness constraints—is
arguably the most notable of these problems to receive widespread attention as a di-
rect result of the development of interior methods (see, e.g., the surveys [65, 99, 101]).
The evident similarity of interior methods to longstanding continuation approaches
(see, e.g., [1, 2]) has been recognized since the early days of modern interior methods
(see, e.g., [71]), but numerous aspects remain to be explored.
As a remarkable bonus, interior methods are playing a growing role in the study
of hard combinatorial problems. Many of the most important problems in discrete
optimization (where the variables are required to be integers) are NP-hard, i.e., they
cannot be solved in polynomial time unless someone favorably resolves the still-open
question of whether P = NP. In the meantime, good approximate solutions are being
found by approximation algorithms—polynomial-time algorithms whose solution is
provably within a certain factor of the optimal solution for the hard problem. A
main ingredient in a successful approximation algorithm is formulation of a convex
relaxation (often a semidefinite program) in which integrality constraints are replaced
by definiteness constraints on associated matrices. An exceptionally clear introduction
to this subject is given in [108].
Almost twenty years after the beginning of the interior-point revolution, there
seems to be no end in sight to new applications of interior methods and new interpre-
tations of the interior-point perspective.
1.2. Classical Barrier Methods. As we have just sketched, classical barrier meth-
ods are closely related to modern interior methods, and we briefly summarize their
history. During the 1960s, the accepted way to solve constrained problems was to
transform them into parameterized unconstrained problems via penalty or barrier
terms. For inequality constraints, a barrier method is motivated by unconstrained

528 ANDERS FORSGREN, PHILIP E. GILL, AND MARGARET H. WRIGHT
minimization of a function combining f and a positively weighted “barrier” that pre-
vents iterates from leaving the feasible region. Penalty methods, in contrast, are based
on minimizing a function that includes f and a positive penalty if evaluated at any
infeasible point.
A large body of beautiful mathematical theory about barrier and penalty func-
tions was developed during the 1960s by Fiacco and McCormick. They also seem to
have introduced the term “interior-point methods” in their seminal book [33, p. 41],
which describes in detail the relationships between minimizers of barrier and penalty
function and solutions of the original constrained problem.
Despite the good features of barrier methods, they were dogged by several con-
cerns. The worry expressed most often in print involved ill-conditioning, after Lootsma
[66] and Murray [73] showed independently in the late 1960s that in general the Hes-
sian of a barrier function becomes increasingly ill-conditioned as the solution is ap-
proached and is singular in the limit. Increasing awareness of this property led to
serious anxiety about the reliability of barrier methods just as other methods were
coming along that seemed to be more efficient in practice without being plagued
by unavoidable ill-conditioning. In particular, augmented Lagrangian and sequen-
tial quadratic programming (SQP) methods (see, for example, [6, 34, 52, 77, 80])
are based directly on the optimality conditions for constrained optimization. Barrier
methods appeared distinctly unappealing by comparison, and almost all researchers
in mainstream optimization lost interest in them.
As described in section 1.1, the dormancy of barrier methods ended in high drama
near the start of the interior-point revolution. An obvious question then needed to
be answered: Are classical barrier methods fundamentally flawed, as once feared?
The answer turns out to be “yes,” but, surprisingly, not because of ill-conditioning.
Classical barrier methods are indeed inefficient—but, by a strange twist of fate, ill-
conditioning, their longtime bugbear, has recently been shown not to be harmful under
circumstances that almost always hold in practice. We explore several interesting
properties, good and bad, of the classical Newton barrier method in section 4.3. An
obvious strategy has been to create interior methods that retain the good properties
of classical barrier methods, yet do not suffer from their defects. The general opinion
today is that primal-dual methods, to be discussed in section 5, offer the greatest
promise for achieving these ends.
It is impossible to cover interior methods for nonlinear optimization thoroughly
in anything less than a large volume. A major goal of this article is thus to show con-
nections between classical and modern ideas and to cover highlights of both theory
and practice; readers interested in learning more about interior-point methods will
find an abundance of papers and books on the subject. Since linear algebra is a spe-
cial interest of the authors, we have devoted extra attention to linear algebraic issues
associated with interior methods. The linear algebra needs of interior methods are
interesting for several reasons. Certain key matrices display increasing ill-conditioning
as the solution is approached, but the ill-conditioning is highly structured. In con-
trast to active-set methods like the simplex method that continually update a set of
constraints temporarily treated as equalities, interior methods typically include all
constraints at every iteration. Hence the matrices arising in interior methods must
somehow reveal, without omitting any constraints, that some constraints are more
important than others. Similarly, two subspaces—the range space of the transposed
Jacobian of the active constraints and the associated null space—strongly affect all
calculations near the solution, but these subspaces are not known explicitly.

INTERIOR METHODS 529
1.3. Statement of the Problem. We concentrate on interior methods for con-
tinuous nonlinear optimization problems of the following form:
(1.1)
minimize
xR
n
f(x)
subject to c
i
(x)=0,i∈E,
c
i
(x) 0,i∈I,
where c(x)isanm-vector of nonlinear constraint functions with ith component c
i
(x),
i =1,...,m, and E and I are nonintersecting index sets. It is assumed throughout
that f and c are twice-continuously differentiable. Any point x satisfying the con-
straints of (1.1) is called a feasible point, and the set of all such points is the feasible
region. We first consider problems containing only inequality constraints (sections 2
through 5) and then turn in sections 6 and 7 to the general form (1.1).
1.4. A Few Words on Coverage and Notation. Since thousands of scientific
papers have been written about interior methods, as already noted we cannot cover
more than a tiny fraction of the field, and it would be equally impractical to cite
all relevant references. We apologize in advance to all those whose favorite topics or
works have not been mentioned here.
Because this is a survey intended for nonexperts, we have included a substantial
amount of background material on optimality conditions in section 2. Readers already
familiar with optimization should skip directly to section 3. Various useful definitions,
lemmas, and miscellaneous results are collected in the appendix.
Finally, because there are not enough letters in the alphabet, especially letters that
are free from previous connotations, we confess to straining at times to find notation
that is clear and precise without being cluttered. To alleviate this dilemma, we
sometimes introduce local abbreviations for the sake of short formulas. For example,
when considering a particular point, say x
, we will sometimes abbreviate quantities
evaluated at x
by adding a superscript and omitting the argument, e.g., we denote
c(x
)byc
. Following common usage in the interior-point literature, if a vector is
denoted by a lowercase letter, the same uppercase letter denotes the diagonal matrix
whose elements are those of the vector, so that V
= diag(v). Finally, e denotes the
vector of all ones whose dimension is determined by the context.
2. Inequality-Constrained Optimization. We begin with problems containing
only inequality constraints:
(2.1) minimize
xR
n
f(x) subject to c(x) 0,
where c(x)isanm-vector of functions {c
i
(x)}, i =1, ...,m, and we assume through-
out that f and {c
i
} are twice-continuously differentiable. The gradient of f is de-
noted by either f(x)org(x), and
2
f(x) denotes the Hessian matrix of second
partial derivatives of f. The gradient and Hessian of c
i
(x) are denoted by c
i
(x)
and
2
c
i
(x). The m × n Jacobian matrix c
(x) of first derivatives of c(x) has rows
{∇c
i
(x)
T
}, and we sometimes (to avoid clutter) use J(x) to denote this Jacobian.
The topic of optimality conditions for nonlinearly constrained optimization can
be complicated and confusing. We present only aspects that will be needed later;
detailed discussions may be found in, for example, [6, 88].
2.1. The KKT Conditions. The terms “KKT point” (standing for “Karush–
Kuhn–Tucker point”) and “KKT conditions” will be used often. In defining these

Figures
Citations
More filters
Journal ArticleDOI

On the implementation of an interior-point filter line-search algorithm for large-scale nonlinear programming

TL;DR: A comprehensive description of the primal-dual interior-point algorithm with a filter line-search method for nonlinear programming is provided, including the feasibility restoration phase for the filter method, second-order corrections, and inertia correction of the KKT matrix.
Journal ArticleDOI

A sparse signal reconstruction perspective for source localization with sensor arrays

TL;DR: This work presents a source localization method based on a sparse representation of sensor measurements with an overcomplete basis composed of samples from the array manifold that has a number of advantages over other source localization techniques, including increased resolution, improved robustness to noise, limitations in data quantity, and correlation of the sources.
Journal ArticleDOI

Control Barrier Function Based Quadratic Programs for Safety Critical Systems

TL;DR: This paper develops a methodology that allows safety conditions—expression as control barrier functions—to be unified with performance objectives—expressed as control Lyapunov functions—in the context of real-time optimization-based controllers.
Book

Convex Optimization & Euclidean Distance Geometry

Jon Dattorro
TL;DR: This book is about convex optimization, convex geometry (with particular attention to distance geometry), geometric problems, and problems that can be transformed into geometrical problems.
Book

Matrix Analysis and Applications

TL;DR: In this article, the authors present the theory, methods and applications of matrix analysis in a new theoretical framework, allowing readers to understand second-order and higher-order matrix analysis.
References
More filters
Book

Numerical Optimization

TL;DR: Numerical Optimization presents a comprehensive and up-to-date description of the most effective methods in continuous optimization, responding to the growing interest in optimization in engineering, science, and business by focusing on the methods that are best suited to practical problems.
Book

Nonlinear Programming

Book

Iterative Solution of Nonlinear Equations in Several Variables

TL;DR: In this article, the authors present a list of basic reference books for convergence of Minimization Methods in linear algebra and linear algebra with a focus on convergence under partial ordering.
Book

Practical Methods of Optimization

TL;DR: The aim of this book is to provide a Discussion of Constrained Optimization and its Applications to Linear Programming and Other Optimization Problems.