An Introduction To Nonlinear Optimization Theory Marius Durea Radu Strugariu
An Introduction To Nonlinear Optimization Theory Marius Durea Radu Strugariu
An Introduction To Nonlinear Optimization Theory Marius Durea Radu Strugariu
An Introduction To Nonlinear Optimization Theory Marius Durea Radu Strugariu
1. An Introduction To Nonlinear Optimization Theory
Marius Durea Radu Strugariu download
https://ebookbell.com/product/an-introduction-to-nonlinear-
optimization-theory-marius-durea-radu-strugariu-51046154
Explore and download more ebooks at ebookbell.com
2. Here are some recommended products that we believe you will be
interested in. You can click the link to download.
An Introduction To Nonlinear Analysis Schechter M
https://ebookbell.com/product/an-introduction-to-nonlinear-analysis-
schechter-m-2047344
An Introduction To Nonlinear Partial Differential Equations 2nd
Edition J David Logan
https://ebookbell.com/product/an-introduction-to-nonlinear-partial-
differential-equations-2nd-edition-j-david-logan-2249850
An Introduction To Nonlinear Functional Analysis And Elliptic Problems
1st Edition Antonio Ambrosetti
https://ebookbell.com/product/an-introduction-to-nonlinear-functional-
analysis-and-elliptic-problems-1st-edition-antonio-ambrosetti-2543308
An Introduction To Nonlinear Finite Element Analysis With Applications
To Heat Transfer Fluid Mechanics And Solid Mechanics 2nd Edition Reddy
https://ebookbell.com/product/an-introduction-to-nonlinear-finite-
element-analysis-with-applications-to-heat-transfer-fluid-mechanics-
and-solid-mechanics-2nd-edition-reddy-34081414
3. An Introduction To Nonlinear Analysis Theory 1st Edition Zdzisaw
Denkowski
https://ebookbell.com/product/an-introduction-to-nonlinear-analysis-
theory-1st-edition-zdzisaw-denkowski-4592578
An Introduction To Nonlinear Analysis And Fixed Point Theory 1st Ed
2018 Hemant Kumar Pathak
https://ebookbell.com/product/an-introduction-to-nonlinear-analysis-
and-fixed-point-theory-1st-ed-2018-hemant-kumar-pathak-7043362
An Introduction To Nonlinear Finite Element Analysis Illustrated
Edition J N Reddy
https://ebookbell.com/product/an-introduction-to-nonlinear-finite-
element-analysis-illustrated-edition-j-n-reddy-981524
Complex Dynamics And Morphogenesis An Introduction To Nonlinear
Science 1st Edition Chaouqi Misbah Auth
https://ebookbell.com/product/complex-dynamics-and-morphogenesis-an-
introduction-to-nonlinear-science-1st-edition-chaouqi-misbah-
auth-5741688
Optima And Equilibria An Introduction To Nonlinear Analysis 2 Ed Aubin
https://ebookbell.com/product/optima-and-equilibria-an-introduction-
to-nonlinear-analysis-2-ed-aubin-11640254
5. Marius Durea and Radu Strugariu
An Introduction to Nonlinear Optimization Theory
7. Marius Durea and Radu Strugariu
An Introduction to
Nonlinear Optimization
Theory
|
Managing Editor: Aleksandra Nowacka-Leverton
Associate Editor: Vicentiu Radulescu
Language Editor: Nick Rogers
12. viii | Preface
gand tangent cone to a set at a point are studied, and we give complete proofs of fun-
damental results, among which Farkas Lemma, Banach Principle of fixed point and
Graves Theorem.
The third chapter is the first one fully dedicated to optimization problems; it
presents in detail the main aspects of the theory for the case of smooth data. We
present general necessary and sufficient optimality conditions of first and second-
order for problems with differentiable cost functions and with geometrical or smooth
functional constraints. We arrive at the famous Karush-Kuhn-Tucker optimality con-
ditions and we investigate several qualification conditions needed in this celebrated
result.
The fourth chapter concerns the case of convex nonsmooth optimization prob-
lems. We introduce here, in compensation for the missing differentiability, the concept
of the subgradient and we deduce, in this setting, necessary optimality conditions in
Fritz John and Karush-Kuhn-Tucker forms.
The fifth chapter generalizes the theory. We work with functions that are neither
differentiable nor convex, but are locally Lipschitz. This is a good setting to present
Clarke and Mordukhovich generalized differentiation calculus, which finally allows
us to arrive, once again, at optimality conditions with similar formulations as in the
previous two chapters.
The sixth chapter is dedicated to the presentation of some basic algorithms for
smooth optimization problems. We show Matlab code that accurately approximate the
solutions of some optimization problems or related nonlinear equations.
The seventh chapter contains more than one hundred exercises and problems
which are organized according to main themes of the book: nonlinear analysis,
smooth optimization, nonsmooth optimization.
In our presentation we used several important monographs as follows: for the-
oretical expositions we mainly used (Zălinescu 1998; Pachpatte 2005; Nocedal and
Wright 2006; Niculescu and Persson 2006; Rădulescu et al. 2009; Mordukhovich 2006;
Hiriart-Urruty 2008; Clarke 2013; Cârjă 2003;), while for examples, problems and ex-
ercises we used (Pedregal 2004; Nocedal and Write, 2006; Isaacson and Keller 1966;
Hiriart-Urruty 2009; Hestenes 1975; Forsgren et al. 2002; Clarke, 1983) and (Bazaraa
et al. 2006). Finally, for the Matlab numerical simulations we used (Quarteroni and
Saleri 2006).
In the case of the Ekeland Variational Principle, which was obtained in 1974 in
the framework of general metric spaces, and whose original proof was based on an
iteration procedure, the simpler proof for the case of finite dimensional vector spaces
we present here was obtained in 1983 by J.-B. Hiriart-Urruty in (Hiriart-Urruty 1983).
The very simple and natural proof of Farkas Lemma that is given in this book is based
on the paper of D. Bartl (Bartl 2012) and on a personal communication to the authors
from C. Zălinescu. The Graves theorem is taken from (Cârjă 2003).
13. | ix
Of course, the main reference for convex analysis is the celebrated R. T. Rockafel-
lar monograph (Rockafellar 1970), but we also used the books (Niculescu and Pers-
son 2006; Zălinescu 1998) and (Zălinescu 2002). For the section concerning the fixed
points for function of real variable, we used several problems presented in (Radulescu
et al. 2009).
For the section dedicated to the generalized Clarke calculus, we used the mono-
graphs (Clarke 1983; Clarke 2013; Rockafellar and Wets 1998). Theorem 5.1.32 is taken
from (Rockafellar 1985). For the second part of Chapter 5, dedicated to Mordukhovich
calculus, we mainly used (Mordukhovich 2006). The calculus rules for the Fréchet
subdifferential of difference of functions, as well as the chain rule for the Fréchet sub-
differential were taken from (Mordukhovich et al. 2006).
Many of the optimization problems given as exercises are taken from (Hiriart-
Urruty 2008) and (Pedregal 2004), but (Hiriart-Urruty 2008) was used as well for some
other theoretical examples such as the second problem from Section 3.4 or the Kan-
torovich inequality. The rather complicated proof of the fact that the Mangasarian Fro-
movitz condition is a qualification condition is taken from (Nocedal and Write 2006)
which is used as well for presenting the sufficient optimality conditions of second
order in Section 3.3. The Hardy and Carleman inequalities correspond to material in
(Pachpatte 2005).
Chapter 6 is dedicated to numerical algorithms. We used the monographs: (Isaac-
son and Keller 1966) (for the convergence of Picard iterations and the Aitken methods),
(Nocedal and Write 2006) (for the Newton and the SQP methods). For the presentation
of the barrier method, we used (Forsgren et al. 2002).
Acknowledgements: We would like to thank Professor Vicenţiu Rădulescu, who
kindly showed us the opportunity to write this book. Then, our thanks are addressed to
dr. Aleksandra Nowacka-Leverton, Managing Editor to De Gruyter Open, for her sup-
port during the preparation of the manuscript, and to the Technical Department of De
Gruyter Open, for their professional contribution to the final form of the monograph.
We also take this opportunity to thank our families for their endless patience during
the many days (including weekends) of work on this book.
01.10.2014 Marius Durea
Iaşi, Romania Radu Strugariu
16. 2 | Preliminaries
The angle between two vectors x, y ∈ Rp
{0} is the value θ ∈ [0, π] given by
cos θ :=
hx, yi
kxk kyk
.
The open (closed) ball and the sphere centered at x ∈ Rp
with radius ε > 0 are given,
respectively, by:
B(x, ε) := {x ∈ Rp
| kx − xk < ε},
D(x, ε) := {x ∈ Rp
| kx − xk ≤ ε},
and
S(x, ε) := {x ∈ Rp
| kx − xk = ε}.
One says that a subset A ⊂ Rp
is bounded if it is contained in an open ball centered
in the origin i.e., if there exists M > 0 such that A ⊂ B(0, M).
A neighborhood of an element x ∈ Rp
is a subset of Rp
which contains an open
ball centered in x. We denote by V(x) the class of all neighborhoods of x. Let us sum-
marize some facts:
– A subset of Rp
is open if it is empty or it is neighborhood for all of its points.
– A subset of Rp
is closed if its complement with respect to Rp
is open.
– An element a is an interior point of the set A ⊂ Rp
if A is a neighborhood of a. We
denote by int A the interior of A (i.e., the set of all interior points of A).
– An element a is an accumulation point (or a limit point) of A ⊂ Rp
if every neigh-
borhood of a has at least one element in common with the set A which is different
from a. We denote by A0
the set of all limit points of A. If a ∈ A A0
, one says that
a is an isolated point of A.
– An element a is an adherent point (or a closure point) of A ⊂ Rp
if every neigh-
borhood of a has at least one element in common with the set A. We will use the
notations cl A and A to denote the closure of A (i.e., the set of all the adherent
points of A).
– A subset of Rp
is compact if it is bounded and closed.
We denote by bd A the set cl A int A = cl A ∩ cl(Rp
A) and we call it the boundary of
A.
Proposition 1.1.1. (i) A subset of Rp
is open if and only if it coincides with its interior.
(ii) A subset of Rp
is closed if and only if it coincides with its closure.
Definition 1.1.2. One says that a function f : N → Rp
is a sequence of elements from
Rp
.
The value of the function f in n ∈ N, f(n), is denoted by xn (or yn, zn, ...), and the
sequence defined by f is denoted by (xn) (respectively, by (yn), (zn), ...).
Definition 1.1.3. A sequence is bounded if the set of its terms is bounded.
17. Rp Space | 3
Definition 1.1.4. One says that (yk) is a subsequence of (xn) if for every k ∈ N, one has
yk = xnk , where by (nk) one denotes a strictly increasing sequence of natural numbers
(i.e., nk < nk+1 for every k ∈ N).
Definition 1.1.5. One says that a sequence (xn) ⊂ Rp
is convergent (or converges) if
there exists x ∈ Rp
such that
∀ V ∈ V(x), ∃nV ∈ N, ∀n ≥ nV : xn ∈ V.
The element x is called the limit of (xn).
If it exists, the limit of a sequence is unique.
We will use the notations xn → x, lim
n→∞
xn = x or, simplified, lim xn = x to formalize
the previous definition.
Proposition 1.1.6. A sequence (xn) is convergent to x ∈ Rp
if and only if
∀ε > 0, ∃nε ∈ N, ∀n ≥ nε : kxn − xk < ε.
Proposition 1.1.7. The sequence (xn) ⊂ Rp
converges to x ∈ Rp
if and only if the
coordinate sequences (xi
n) converge (in R) to xi
for every i ∈ 1, p.
Proposition 1.1.8. A sequence is convergent to x ∈ Rp
if and only if all of its subse-
quences are convergent to x.
Proposition 1.1.9. Every convergent sequence is bounded.
Proposition 1.1.10 (Characterization of the closure points using sequences). Con-
sider A ⊂ Rp
. A point x ∈ Rp
is a closure point of A if and only if there exists a sequence
(xn) ⊂ A such that xn → x.
Proposition 1.1.11. The set A ⊂ Rp
is closed if and only if every convergent sequence
from A has its limit in A.
Proposition 1.1.12. The set A ⊂ Rp
is compact if and only if every sequence from A has
a subsequence which converges to a point of A.
Theorem 1.1.13 (Cesàro Lemma). Every bounded sequence contains a convergent sub-
sequence.
Definition 1.1.14. One says that (xn) ⊂ Rp
is a Cauchy sequence or a fundamental
sequence if
∀ε > 0, ∃nε ∈ N, ∀n, m ≥ nε : kxn − xmk < ε.
The above definition can be reformulated as follows: (xn) is a Cauchy sequence if:
18. 4 | Preliminaries
∀ε > 0, ∃nε ∈ N, ∀n ≥ nε, ∀p ∈ N : kxn+p − xnk < ε.
Theorem 1.1.15 (Cauchy). The space Rp
is complete, i.e., a sequence from Rp
is con-
vergent if and only if it is a Cauchy sequence.
The next results are specific to the case of real sequences.
Definition 1.1.16. One says that a sequence (xn) of real numbers is increasing (strictly
increasing, decreasing, strictly decreasing) if for every n ∈ N, xn+1 ≥ xn (xn+1 > xn,
xn+1 ≤ xn, xn+1 < xn). If (xn) is either increasing or decreasing, then it is called monotone.
Let R := R ∪ {−∞, +∞} be the set of extended real numbers. A neighborhood of +∞
is a subset of R which contains an interval of the form (x, +∞], where x ∈ R. The
neighborhoods of −∞ are defined in a similar manner.
Definition 1.1.17. (i) One says that the sequence (xn) ⊂ R has the limit equal to +∞ if
∀V ∈ V(+∞), ∃nV ∈ N, ∀n ≥ nV : xn ∈ V.
(ii) One says that the sequence (xn) ⊂ R has the limit equal to −∞ if
∀V ∈ V(−∞), ∃nV ∈ N, ∀n ≥ nV : xn ∈ V.
Proposition 1.1.18. (i) A sequence (xn) ⊂ R has the limit equal to +∞ if and only if
∀A > 0, ∃nA ∈ N, ∀n ≥ nA : xn > A.
(ii) A sequence (xn) ⊂ R has the limit equal to −∞ if and only if
∀A > 0, ∃nA ∈ N, ∀n ≥ nA : xn < −A.
Proposition 1.1.19. Let (xn), (yn), (zn) be sequences of real numbers, x, y ∈ R and n0 ∈
N. Then:
(i) (Passing to the limit in inequalities) if xn → x, yn → y and xn ≤ yn for every
n ≥ n0, then x ≤ y;
(ii) (The boundedness criterion) if |xn − x| ≤ yn for every n ≥ n0, and yn → 0, then
xn → x;
(iii) if xn ≥ yn for every n ≥ n0, and yn → +∞, then xn → +∞;
(iv) if xn ≥ yn for every n ≥ n0, and xn → −∞, then yn → −∞;
(v) if (xn) is bounded and yn → 0, then xnyn → 0;
(vi) if xn ≤ yn ≤ zn for every n ≥ n0, and xn → x, zn → x, then yn → x;
(vii) xn → 0 ⇔ |xn| → 0 ⇔ x2
n → 0.
We now present some fundamental results in the theory of real sequences.
19. Limits of Functions and Continuity | 5
Theorem 1.1.20. Every monotone real sequence has its limit in R. Moreover, if the se-
quence is bounded, then it is convergent, as follows: if it is increasing, then its limit is the
supremum of the set of its terms, and if it is decreasing, the limit is the infimum of the set
of its terms. If it is unbounded, then its limit is either +∞ if the sequence is increasing, or
−∞ if the sequence is decreasing.
Theorem 1.1.21 (Weierstrass theorem for sequences). If (xn) is a bounded and mono-
tone sequence of real numbers, then (xn) is convergent.
Definition 1.1.22. Let (xn)n≥0 be a sequence of real numbers. An element x ∈ R is called
a limit point of (xn) if there exists a subsequence (xnk ) of (xn) such that x = lim
k→∞
xnk .
We finalize this section with two useful convergence criteria.
Proposition 1.1.23. Let (xn) be a sequence of strictly positive real numbers such that
there exists lim
xn+1
xn
= x. If x < 1, then xn → 0, and if x > 1, then xn → +∞.
Proposition 1.1.24 (Stolz-Cesàro Criterion). Let (xn) and (yn) be real sequences such
that (yn) is strictly increasing and its limit is equal to +∞. If there exists lim
xn+1 − xn
yn+1 − yn
=
x ∈ R, then lim
xn
yn
exists and is equal to x.
1.2 Limits of Functions and Continuity
In this section, we expand on some issues related to the concepts of limit and conti-
nuity for functions. Let p, q ∈ N*
.
Definition 1.2.1. Let f : A → Rq
, A ⊂ Rp
and a ∈ A0
. One says that the element l ∈ Rq
is the limit of the function f at a, if for every V ∈ V (l) , there exists U ∈ V (a) such that
if x ∈ U ∩ A, x ≠ a, then f (x) ∈ V. We will denote this situation by lim
x→a
f (x) = l.
Theorem 1.2.2. Let f : A → Rq
, A ⊂ Rp
and a ∈ A0
. The next assertions are equiva-
lent:
(i) lim
x→a
f (x) = l;
(ii) for every B (l, ε) ⊂ Rq
, there exists B (a, δ) ⊂ Rp
such that if x ∈ B (a, δ) ∩ A,
x ≠ a, then f (x) ∈ B (l, ε) ;
(iii) for every ε > 0, there exists δ > 0, such that if kx − ak < δ, x ∈ A, x ≠ a, then
kf (x) − lk < ε;
(iv) for every ε > 0, there exists δ > 0, such that if |xi − ai| < δ for every i ∈ 1, p,
where x = (x1, x2, .., xp) ∈ A, a = (a1, a2, .., ap) , x ≠ a, then kf (x) − lk < ε;
(v) for every sequence (xn) ⊂ A {a} , xn → a implies that f (xn) → l.
20. 6 | Preliminaries
Theorem 1.2.3. Let f : A → Rq
, A ⊂ Rp
, l ∈ Rq
and a ∈ A0
. If the function f has the
limit l at a, then this limit is unique.
Remark 1.2.4. If there exist two sequences x0
n
, x00
n
⊂ A {a} , x0
n → a, x00
n → a
such that f(x0
n) → l0
, f(x00
n ) → l00
and l0
≠ l00
, then the limit of the function f at a ∈ A0
does not exist.
Theorem 1.2.5. Let f : A → Rq
, A ⊂ Rp
, f = (f1, f2, ..., fq) and a ∈ A0
. Then f has
the limit l = (l1, l2, ..., lq) ∈ Rq
at a if and only if there exists lim
x→a
fi (x) = li, for every
i ∈ 1, q.
Definition 1.2.6. Let a ∈ R, A ⊂ R and denote As = A ∩(−∞, a], Ad = A ∩[a, ∞). One
says that the element a is a left (right) accumulation point for A, if it is an accumulation
point for As (Ad, respectively). We will denote the set of left (right) accumulation points
of A by A0
s (A0
d, respectively).
Definition 1.2.7. Let f : A → Rq
, A ⊂ R and a be a left (right) accumulation point of
A. One says that the element l ∈ Rq
is the left-hand (right-hand) limit of the function f
in a if for every neighborhood V ∈ V (l) there exists U ∈ V(a), such that if x ∈ U ∩ As
(x ∈ U ∩ Ad, respectively), x ≠ a, then f (x) ∈ V. In this case we will write lim
x→a,xa
f (x) =
l, or lim
x→a−
f (x) = l, or lim
x↑a
f (x) = l ( lim
x→a,xa
f (x) = l, or lim
x→a+
f (x) = l, or lim
x↓a
f (x) = l,
respectively).
Theorem 1.2.8. Let I ⊂ R be an open interval, f : I → Rq
, and a ∈ I. Then there exists
lim
x→a
f (x) = l if and only if the left-hand and the right-hand limits of f at a exist and they
are equal. In this case, all three limits are equal:
lim
x→a−
f (x) = lim
x→a+
f (x) = l.
A well-known result says that the monotone real functions admit lateral limits at every
accumulation point of their domains.
Theorem 1.2.9 (The boundedness criterion). Let f : A → Rq
, g : A → R, A ⊂ Rp
and a ∈ A0
. If there exist l ∈ Rq
and U ∈ V (a) such that kf (x) − lk ≤ |g (x)| for every
x ∈ U {a} , and lim
x→a
g (x) = 0, then there exists lim
x→a
f (x) = l.
Theorem 1.2.10. Let f , g : A ⊂ Rp
→ Rq
, and a ∈ A0
. If lim
x→a
f (x) = 0 and there exists
U ∈ V (a) such that g is bounded on U, then there exists the limit lim
x→a
f (x) g (x) = 0.
Theorem 1.2.11. Let f : A ⊂ Rp
→ Rq
, and a ∈ A0
. If there exists lim
x→a
f (x) = l, l 0
(l 0), then there exists U ∈ V (a) such that for every x ∈ U ∩A, x ≠ a, one has f (x) 0
(respectively, f (x) 0).
21. Limits of Functions and Continuity | 7
Theorem 1.2.12. Let f : A ⊂ Rp
→ Rq
, and a ∈ A0
. If there exists lim
x→a
f (x) = l, then
there exists U ∈ V (a) such that f is bounded on U (i.e., there exists M 0 such that for
every x ∈ U, one has kf (x)k ≤ M).
Definition 1.2.13. Let f : A ⊂ Rp
→ R and a ∈ A0
. One says that the function f
has the limit equal to +∞ (respectively, −∞) at a, if for every V ∈ V (+∞) (respectively,
V ∈ V (−∞)), there exists U ∈ V (a) such that for every x ∈ U ∩ A, x ≠ a, one has
f (x) ∈ V. In this case, we will write lim
x→a
f (x) = +∞ (respectively, lim
x→a
f (x) = −∞).
Theorem 1.2.14. Let f : A ⊂ Rp
→ R and a ∈ A0
. Then there exists lim
x→a
f (x) = +∞
(respectively, lim
x→a
f (x) = −∞) if and only if for every ε 0, there exists δ 0, such that
if kx − ak δ, x ∈ A, x ≠ a, one has f (x) ε (respectively, f (x) −ε).
Definition 1.2.15. Let f : A ⊂ R → Rq
, such that +∞ (respectively, −∞) is an accumu-
lation point of A. One says that the element l ∈ Rq
is the limit of f at +∞ (respectively,
−∞), if for every V ∈ V (l) , there exists U ∈ V (+∞) (respectively, U ∈ V (−∞)) such
that for every x ∈ U ∩ A, one has f (x) ∈ V. In this case, we will write lim
x→+∞
f (x) = l
(respectively, lim
x→−∞
f (x) = l).
Theorem 1.2.16. Let f : A ⊂ R → Rq
, such that +∞ (respectively, −∞) is an accumu-
lation point of A. Then there exists lim
x→+∞
f (x) = l (respectively, lim
x→−∞
f (x) = l) if and only
if for every ε 0, there exists δ 0, such that if x δ (respectively, x −δ), x ∈ A, one
has kf (x) − lk ε.
Definition 1.2.17. Let f : A ⊂ Rp
→ Rq
, and a ∈ A. One says that the function f
is continuous at a if for every V ∈ V(f(a)), there exists U ∈ V(a) such that for every
x ∈ U ∩ A, one has f(x) ∈ V.
If the function f is not continuous at a, one says that f is discontinuous at a, or that a
is a discontinuity point of the function f.
Theorem 1.2.18. Let f : A ⊂ Rp
→ Rq
, and a ∈ A0
∩ A. The function f is continuous
at a if and only if lim
x→a
f(x) = f(a). If a is an isolated point of A, then f is continuous at a.
Theorem 1.2.19. Let f : A ⊂ Rp
→ Rq
, and a ∈ A. The next assertions are equivalent:
(i) f is continuous at a;
(ii) (ε − δ characterization) for every ε 0, there exists δ 0, such that if kx − ak
δ, x ∈ A, then f(x) − f(a) ε;
(iii) (sequential characterization) for every (xn) ⊂ A, xn → a, one has f(xn) →
f(a).
22. 8 | Preliminaries
Theorem 1.2.20. The image of a compact set through a continuous function is a com-
pact set.
Theorem 1.2.21 (Weierstrass Theorem). Let K be a compact subset of Rp
. If f : K → R
is a continuous function, then f is bounded and it attains its extreme values on the set K
(i.e., there exist a, b ∈ K, such that sup
x∈K
f(x) = f(a) and inf
x∈K
f(x) = f(b)).
Definition 1.2.22. Let f : D ⊂ Rp
→ Rq
. One says that the function f is uniformly
continuous on the set D if for every ε 0, there exists δ 0, such that for every x0
, x00
∈ D
with x0
− x00
δ, one has f(x0
) − f(x00
) ε.
Remark 1.2.23. Every function which is uniformly continuous on D is continuous on D,
i.e., it is continuous at every point of D.
Theorem 1.2.24 (Cantor Theorem). Every function which is continuous on a compact
set K ⊂ Rp
and takes values in Rq
is uniformly continuous on K.
Definition 1.2.25. Let L ≥ 0 be a real number. One says that a function f : A ⊂ Rp
→
Rq
is Lipschitz on A with modulus L, or L−Lipschitz on A, if f(x) − f(y) ≤ L kx − yk,
for every x, y ∈ A.
Proposition 1.2.26. Every Lipschitz function on A ⊂ Rp
is uniformly continuous on A.
Theorem 1.2.27. Let I ⊂ R be an interval. If f : I → R is injective and continuous, then
f is strictly monotone on I.
Definition 1.2.28. Let I ⊂ R be an interval. One says that the function f : I → R
has the Darboux property if for every a, b ∈ I, a b and every λ ∈ (f(a), f(b)) or
λ ∈ (f(b), f(a)), there exists cλ ∈ (a, b) such that f(cλ) = λ.
Theorem 1.2.29. Let I ⊂ R be an interval. If the function f : I → R has the Darboux
property and there exist a, b ∈ I, a b, such that f(a)f(b) 0, then the equation
f(x) = 0 has at least one solution in (a, b).
Theorem 1.2.30. Let I ⊂ R be an interval. The function f : I → R has the Darboux
property if and only if for every interval J ⊂ I, f (J) is an interval.
Theorem 1.2.31. Let I ⊂ R be an interval. If f : I → R is continuous, then f has the
Darboux property.
23. Limits of Functions and Continuity | 9
Recall that every linear operator T : Rp
→ Rq
is continuous. For such a map, one uses
the constant
kTk := inf{M 0 | kTxk ≤ M kxk , ∀x ∈ Rp
}
= sup
kTxk | x ∈ D(0, 1) .
The mapping T 7→ kTk satisfies the axioms of a norm, therefore it is called the norm of
the operator T. Consequently, the set of linear operators from Rp
to Rq
is a real normed
vector space, with respect to the usual algebraic operations and to the norm previously
defined. This space is denoted by L(Rp
, Rq
) and can be isomorphically identified with
Rpq
. Every operator T ∈ L(Rp
, Rq
) can be naturally associated with a q × p matrix,
denoted by AT = (aji)j∈1,q,i∈1,p, as follows: if (ei)i∈1,p and (e0
i)i∈1,q are the canonical
bases of the spaces Rp
and Rq
, respectively, then (aji)j∈1,q,i∈1,p are the coordinates of
the expressions of the images of the elements (ei)i∈1,p through T with respect to the
basis (e0
i)i∈1,q, i.e.,
T(ei) =
q
X
j=1
ajie0
j , ∀i ∈ 1, p.
Consequently, T 7→ AT is an isomorphism of linear spaces between L(Rp
, Rq
) and the
space of real q × p matrices. Also, for every x ∈ Rp
:
T(x) = (AT xt
)t
.
Moreover, for every x ∈ Rp
and y ∈ Rq
, one has that
D
(AT xt
)t
, y
E
=
D
x, (At
T yt
)t
E
.
If A is a q×p matrix, then the linear operator associated with A is surjective if and only
if the map associated with At
is injective.
Recall also that if T : Rp
→ Rq
is a linear operator, then its kernel,
Ker(T) := {x ∈ Rp
| T(x) = 0},
is a linear subspace of Rp
, and its image,
Im(T) := {T(x) | x ∈ Rp
},
is a linear subspace of Rq
. Moreover,
p = dim(Ker(T)) + dim(Im(T)),
where by dim we denote the algebraic dimension.
From the theory of linear algebra, one knows that if A is a symmetric square ma-
trix of order p, then its eigenvalues are real and, moreover, there exists an orthogonal
24. 10 | Preliminaries
matrix B (i.e., BBt
= Bt
B = I) such that Bt
AB is the diagonal matrix having the eigen-
values on its main diagonal. Recall that, as usual, I denotes the identity matrix.
One says that a matrix A as above is positive semidefinite if (Axt
)t
, x ≥ 0 for
every x ∈ Rp
, and positive definite if (Axt
)t
, x 0 for every x ∈ Rp
{0}. Actually,
A is positive definite if and only if it is positive semidefinite and invertible.
We end this section by mentioning the celebrated result of Hahn-Banach. Recall
that a function f : Rp
→ R is called sublinear if it is positive homogeneous (i.e.,
f(αx) = αf(x) for all α ≥ 0 and x ∈ Rp
) and subadditive (i.e., f(x + y) ≤ f(x) + f(y) for
all x, y ∈ Rp
).
Theorem 1.2.32 (Hahn-Banach). Let X be a linear subspace of Rp
, χ : Rp
→ R be a
sublinear function, and φ0 : X → R be a linear function. If φ0(x) ≤ χ(x) for every x ∈ X,
then there exists a linear function φ : Rp
→ R such that φ X = φ0 and φ(x) ≤ χ(x) for
every x ∈ Rp
.
1.3 Differentiability
Definition 1.3.1. Let f : D ⊂ Rp
→ Rq
and a ∈ int D. One says that f is Fréchet
differentiable (or, simply, differentiable) at a if there exists a linear operator denoted by
∇f(a) : Rp
→ Rq
such that
lim
h→0
f(a + h) − f(a) − ∇f(a)(h)
khk
= lim
x→a
f(x) − f(a) − ∇f(a)(x − a)
kx − ak
= 0.
The map ∇f(a) is called the Fréchet differential of the function f at a.
The previous relation is equivalent to the following conditions:
∀ε 0, ∃δ 0, ∀x ∈ B(a, δ) : f(x) − f(a) − ∇f(a)(x − a) ≤ ε kx − ak ;
∃α : D − {a} → Rq
: lim
h→0
α(h) = α(0) = 0,
f(a + h) = f(a) + ∇f(a)(h) + khk α(h), ∀h ∈ D − {a}.
One says that f : D ⊂ Rp
→ Rq
is of class C1
on the open set D if f is Fréchet
differentiable on D and ∇f is continuous on D. Obviously, f can be written as f =
(f1, f2, ..., fq), where fi : Rp
→ R, i ∈ 1, q and, in general, the map ∇f(a) ∈ L(Rp
, Rq
)
will be identified to the q × p matrix
∂f1
∂x1
(a)
∂f1
∂x2
(a) · · ·
∂f1
∂xp
(a)
∂f2
∂x1
(a)
∂f2
∂x2
(a) · · ·
∂f2
∂xp
(a)
.
.
.
.
.
.
...
.
.
.
∂fq
∂x1
(a)
∂fq
∂x2
(a) · · ·
∂fq
∂xp
(a)
,
25. Differentiability | 11
called the Jacobian matrix of f at the point a, where
∂fi
∂xj
(a) is the partial derivative of
the function fi with respect to the variable xj
at a.
We will subsequently refer several times to the Jacobian matrix instead of the dif-
ferential. Based on a general result, if f : D ⊂ Rp
→ Rp
and a ∈ int D, ∇f(a) is an
isomorphism of Rp
if and only if the Jacobian matrix of f at a is invertible.
The next calculus rules hold.
– Let f : Rp
→ Rq
be an affine function, i.e., it takes the form f(x) := g(x) + u for
every x ∈ Rp
, where g : Rp
→ Rq
is linear, and u ∈ Rq
. Then for every x ∈ Rp
,
∇f(x) = g.
– Let f : Rp
→ R be of the form f(x) =
1
2
(Axt
)t
, x +hb, xi , where A is a symmetric
square matrix of order p, and b ∈ Rp
. Then for every x ∈ Rp
, ∇f(x) = (Axt
)t
+ b.
– Let D ⊂ Rp
, E ⊂ Rq
, x ∈ int D, y ∈ int E and f , g : D → Rq
, φ : D → R,
h : E → Rk
.
– If f , g are differentiable at x, and α, β ∈ R, then the function αf + βg is differ-
entiable at x and
∇(αf + βg)(x) = α∇f(x) + β∇g(x).
– If f , φ are differentiable at x, then φf is differentiable at x at
∇(φf)(x) = φ(x)∇f(x) + f(x)∇φ(x),
where f(x)∇φ(x)
(x) := ∇φ(x)(x) · f(x).
– (Chain rule) If f(D) ⊂ E, y = f(x), f is differentiable at x and h is differentiable
at y, then h ◦ f is differentiable at x and
∇(h ◦ f)(x) = ∇h(y) ◦ ∇f(x).
A casewhich deservesspecial attention is p = 1. In thiscase one saysthat f is derivable
at a if there exists
lim
h→0
f(a + h) − f(a)
h
∈ Rq
. (1.3.1)
One denotes this limit by f0
(a) and it is called the derivative of f at a.
Proposition 1.3.2. Let f : D ⊂ R → Rq
and a ∈ int D. The next assertions are equiva-
lent:
(i) f is derivable at a;
(ii) f is Fréchet differentiable at a.
In every one of these cases, ∇f(a)(x) = xf0
(a) for every x ∈ R.
Let r ∈ N*
. If f : D ⊂ Rp
× Rq
→ Rr
, and (a, b) ∈ int D is fixed, one defines
D1 := {x ∈ Rp
| (x, b) ∈ D} and f1 : D1 → Rr
, f1(x) := f(x, b). One says that f is
Fréchet differentiable with respect to x at a if f1 is Fréchet differentiable at a, and in
26. 12 | Preliminaries
this case the differential is denoted by ∇xf(a, b). If f is differentiable at (a, b), then f
is differentiable with respect to x and y at a and b, respectively, and
∇xf(a, b) = ∇f(a, b)(·, 0), ∇yf(a, b) = ∇f(a, b)(0, ·).
In the general case, one says that f : D ⊂ Rp
→ Rq
is twice Fréchet differentiable
at a ∈ int D if f is Fréchet differentiable on a neighborhood V ⊂ D of a and ∇f :
V → L(Rp
, Rq
) is Fréchet differentiable at a, i.e., there exists a functional denoted by
∇2
f(a), from the space L2
(Rp
, Rq
) := L(Rp
, L(Rp
, Rq
)), and α : D − {a} → L(Rp
, Rq
),
such that limh→0 α(h) = α(0) = 0 and for every h ∈ D − {a}, one has
∇f(a + h) = ∇f(a) + ∇2
f(a)(h, ·) + khk α(h).
Recall that the space L2
(Rp
, Rq
) mentioned above can be identified with the space
of bilinear maps from Rp
× Rp
to Rq
.
One says that f is of class C2
on the open set D if it is twice Fréchet differentiable
on D and ∇2
f : D → L2
(Rp
, Rq
) is continuous.
Theorem 1.3.3. Let f : D ⊂ Rp
→ Rq
and a ∈ int D. If f is twice Fréchet differentiable
at a, then ∇2
f(a) is a symmetric bilinear map.
In the case when q = 1, the map ∇2
f(a) is defined by the symmetric square matrix
H(a) =
∂2
f
∂xi∂xj
(a)
i,j∈1,p
, which is called the Hessian matrix of f at a. Moreover,
D
H(a)ut
t
, v
E
= ∇2
f(a)(u, v) for every u, v ∈ Rp
, i.e.,
∇2
f(a)(u, v) =
n
X
i,j=1
∂2
f
∂xi∂xj
(a)uivj.
If a, b ∈ Rp
, one defines the closed and the open line segments between a and b
as follows:
[a, b] := {αa + (1 − α)b | α ∈ [0, 1]},
(a, b) := {αa + (1 − α)b | α ∈ (0, 1)}.
Theorem 1.3.4 (Lagrange and Taylor Theorems). Let U ⊂ Rp
be an open set, f : U →
R and a, b ∈ U with [a, b] ⊂ U. If f is of class C1
on U, then there exists c ∈ (a, b) such
that
f(b) = f(a) + ∇f(c)(b − a).
If f is of class C2
on U, then there exists c ∈ (a, b) such that
f(b) = f(a) + ∇f(a)(b − a) +
1
2
∇2
f(c)(b − a, b − a).
27. Differentiability | 13
Theorem 1.3.5 (Implicit Function Theorem). Let D ⊂ Rp
× Rq
be an open set, h : D →
Rq
be a function and x ∈ Rp
, y ∈ Rq
be such that:
(i) h(x, y) = 0;
(ii) the function h is of class C1
on D;
(iii) ∇yh(x, y) is invertible.
Then there exist two neighborhoods U and V of x and y, respectively, and a unique
continuous function φ : U → V such that:
(a) h(x, φ(x)) = 0 for every x ∈ U;
(b) if (x, y) ∈ U × V and h(x, y) = 0, then y = φ(x);
(c) φ is differentiable on U and
∇φ(x) = −[∇yh(x, φ(x))]−1
∇xh(x, φ(x)), ∀x ∈ U.
Some fundamental results from the theory of differentiability of the real functions are
briefly given at the end of this section.
In the case p = q = 1 one can apply Proposition 1.3.2. It is also sensible to speak of
the existence of the derivative at points of the domain which are accumulation points
of it: consider the limit from the relation (1.3.1) at accumulation points. Moreover, as in
the case of the lateral limits, one can speak about the left and right-hand derivatives,
by considering the lateral limits in the expression from relation (1.3.1). When they ex-
ist, we will call these limits the left, and the right-hand derivatives of the function f at
a and we will denote them by f0
−(a) and f0
+(a), respectively.
Definition 1.3.6. Let A ⊂ R and f : A → R. One says that a ∈ A is a local minimum
(maximum) point for f if there exists a neighborhood V of a such that f(a) ≤ f(x) (re-
spectively, f(a) ≥ f(x)), for every x ∈ A ∩ V. One says that a point is a local extremum if
it is a local minimum or a local maximum.
Theorem 1.3.7 (Fermat Theorem). Let I ⊂ R be an interval and a ∈ int I. If f : I → R
is derivable at a, and a is a local extremum point for f , then f0
(a) = 0.
Theorem 1.3.8 (Rolle Theorem). Let a, b ∈ R, a b, and f : [a, b] → R be a function
which is continuous on [a, b], derivable on (a, b), and satisfies f(a) = f(b). Then there
exists c ∈ (a, b) such that f0
(c) = 0.
Theorem 1.3.9 (Lagrange Theorem). Let a, b ∈ R, a b, and f : [a, b] → R be a
function which is continuous on [a, b], derivable on (a, b). Then there exists c ∈ (a, b)
such that f(b) − f(a) = f0
(c)(b − a).
Proposition 1.3.10. Let I ⊂ R be an interval and f : I → R be derivable on I.
(i) If f0
(x) = 0 for every x ∈ I, then f is constant on I.
28. 14 | Preliminaries
(ii) If f0
(x) 0 (respectively, if f0
(x) ≥ 0) for every x ∈ I, then f is strictly increasing
(respectively, it is increasing) on I.
(iii) If f0
(x) 0 (respectively, if f0
(x) ≤ 0), for every x ∈ I, then f is strictly decreasing
(respectively, it is decreasing) on I.
Theorem 1.3.11 (Rolle Sequence). Let I ⊂ R be an interval and f : I → R be a deriv-
able function. If x1, x2 ∈ I, x1 x2 are consecutive roots of the derivative f0
(i.e.,
f0
(x1) = 0, f0
(x2) = 0 and f0
(x) ≠ 0 for any x ∈ (x1, x2)) then:
(i) if f(x1)f(x2) 0, the equation f(x) = 0 has exactly one root in the inter-
val (x1, x2);
(ii) if f(x1)f(x2) 0, the equation f(x) = 0 has no roots in the interval (x1, x2);
(iii) if f(x1) = 0 or f(x2) = 0, then x1 or x2 is a multiple root of the equation f(x) = 0
and this equation has no other root in the interval (x1, x2).
Theorem 1.3.12 (Cauchy Rule). Let I ⊂ R be an interval and f , g : I → R, a ∈ I, which
satisfy:
(i) f(a) = g(a) = 0;
(ii) f , g are derivable at a;
(iii) g0
(a) ≠ 0.
Then there exists V ∈ V(a) such that g(x) ≠ 0, for any x ∈ V {a} and
lim
x→a
f(x)
g(x)
=
f0
(a)
g0(a)
.
Theorem 1.3.13 (L’Hôpital Rule). Let f , g : (a, b) → R, where −∞ ≤ a b ≤ ∞. If:
(i) f , g are derivable on (a, b) with g0
≠ 0 on (a, b);
(ii) there exists lim
x→a
xa
f0
(x)
g0(x)
= L ∈ R;
(iii) lim
x→a
xa
f(x) = lim
x→a
xa
g(x) = 0 or
(iii)’ lim
x→a
xa
g(x) = ∞,
then there exists lim
x→a
f(x)
g(x)
= L.
Theorem 1.3.14. Let I ⊂ R be an open interval, f : I → R be a n−times derivable
function at a ∈ I, (n ∈ N, n ≥ 2), such that
f0
(a) = 0, f00
(a) = 0, ..., f(n−1)
(a) = 0, f(n)
(a) ≠ 0.
(i) If n is even, then a is an extremum point, more precisely: a local maximum if
f(n)
(a) 0, and a local minimum if f(n)
(a) 0.
(ii) If n is odd, then a is not an extremum point.
29. The Riemann Integral | 15
1.4 The Riemann Integral
At the end of this chapter we discuss the main aspects concerning the Riemann inte-
gral. Let a, b ∈ R, a b.
Definition 1.4.1. (i) A partition of the interval [a, b] is a finite set of real numbers
x0, x1, ..., xn (n ∈ N*
), denoted by ∆, such that
a = x0 x1 ... xn−1 xn = b.
(ii) The norm of the partition ∆ is the number
k∆k := max{xi − xi−1 | i ∈ 1, n}.
(iii) A tagged partition of the interval [a, b] is a partition ∆, together with a finite set
of real numbers Ξ := {ξi | i ∈ 1, n}, such that ξi ∈ [xi−1, xi] for any i ∈ 1, n. The set Ξ
is called the intermediate points system associated to ∆.
(iv) Let f : [a, b] → R be a function. The Riemann sum associated to a tagged
partition of the interval [a, b] is
S(f , ∆, Ξ) :=
n
X
i=1
f(ξi)(xi − xi−1).
Definition 1.4.2. Let f : [a, b] → R be a function. One says that f is Riemann integrable
on [a, b] if there exists I ∈ R such that for every ε 0, there exists δ 0 such that for
any partition ∆ of the interval [a, b] with the property k∆k δ, and for any intermediate
points system Ξ associated to ∆, the next inequality holds:
S(f , ∆, Ξ) − I ε.
The real number I from the previous definition, which is unique, is called the Riemann
integral of f an [a, b] and is denoted by
b
Z
a
f(x)dx.
Theorem 1.4.3. Any function which is Riemann integrable on [a, b] is bounded on
[a, b].
Definition 1.4.4. Let f : [a, b] → R be a function. One says that a function F : [a, b] →
R is an antiderivative (or, equivalently, a primitive integral, or an indefinite integral) of
f on [a, b] if F is derivable on [a, b] and F0
(x) = f(x) for any x ∈ [a, b].
If an antiderivative exists for a given function, then infinitely many antiderivatives
exist for that function and the difference of any two such antiderivatives is a constant.
The next result is sometimes called the fundamental theorem of calculus.
30. 16 | Preliminaries
Theorem 1.4.5 (Leibniz-Newton). If f : [a, b] → R is Riemann integrable on the inter-
val [a, b] and it admits an antiderivative F on [a, b], then
b
Z
a
f(x)dx = F(b) − F(a).
Continuous functions satisfy both hypotheses of the preceding theorem.
Theorem 1.4.6. If f : [a, b] → R is continuous on [a, b], then f is Riemann integrable
on [a, b] and it admits antiderivatives on [a, b].
Theorem 1.4.7. If f : [a, b] → R is bounded and has a finite set of discontinuity points,
then f is Riemann integrable on [a, b]. Every function which is monotone on [a, b] is
Riemann integrable on [a, b].
We present now the main properties of the Riemann integral.
Theorem 1.4.8. (i) If f , g : [a, b] → R are Riemann integrable on [a, b], and α, β ∈ R,
then αf + βg is Riemann integrable on [a, b] and
b
Z
a
(αf(x) + βg(x))dx = α
b
Z
a
f(x)dx + β
b
Z
a
g(x)dx.
(ii) If f : [a, b] → R is Riemann integrable on [a, b], and m ≤ f(x) ≤ M for every
x ∈ [a, b] (m, M ∈ R), then
m(b − a) ≤
b
Z
a
f(x)dx ≤ M(b − a).
In particular, if f(x) ≥ 0 for every x ∈ [a, b], then
b
Z
a
f(x)dx ≥ 0,
and if f , g : [a, b] → R are Riemann integrable and f(x) ≤ g(x) for every x ∈ [a, b], then
b
Z
a
f(x)dx ≤
b
Z
a
g(x)dx.
(iii) If f : [a, b] → R is Riemann integrable on [a, b], then |f| is Riemann integrable
on [a, b].
(iv) If f , g : [a, b] → R are Riemann integrable [a, b], then f · g is Riemann inte-
grable on [a, b].
31. The Riemann Integral | 17
Theorem 1.4.9. (i) If f : [a, b] → R is Riemann integrable on [a, b], then f is Riemann
integrable on every subinterval of [a, b].
(ii) If c ∈ (a, b) and f is Riemann integrable on [a, c] and on [c, b], then f is Rie-
mann integrable on [a, b] and
b
Z
a
f(x)dx =
c
Z
a
f(x)dx +
b
Z
c
f(x)dx.
Theorem 1.4.10. Let f : [a, b] → R be a function, and f*
: [a, b] → R be another
function which coincides with f on [a, b], except on a finite set of points. If f*
is Riemann
integrable on [a, b], then f is Riemann integrable on [a, b] and
b
Z
a
f(x)dx =
b
Z
a
f*
(x)dx.
Theorem 1.4.11 (integration by parts). If f , g : [ a, b ] → R are C1
functions, then
b
Z
a
f(x)g0
(x) dx = f(x)g(x)| b
a −
b
Z
a
f0
(x)g(x) dx.
Theorem 1.4.12 (change of variable). Let φ : [a, b] → [c, d] be a C1
function, and let
f : [c, d] → R be a continuous function. Then
b
Z
a
f(φ(t)) · φ0
(t) dt =
φ(b)
Z
φ(a)
f(x) dx. (1.4.1)
We end this section by the next multidimensional variant of Taylor Theorem. In what
follows, the equality is understood on components (i.e., for every function fi : Rp
→
R, i = 1, p, where f = (f1, ..., fp)).
Theorem 1.4.13. Suppose f : Rp
→ Rp
is continuously differentiable on some convex
open set D and that x, x + y ∈ D. Then there is t ∈ (0, 1) such that
f(x + y) = f(x) +
1
Z
0
∇f(x + ty)(y) dt.
33. Convex Sets and Cones | 19
one can deduce that u ∈ C. For the converse implication, suppose that C is a cone
which satisfies C + C = C. Fix α ∈ (0, 1) and c1, c2 ∈ C. Then, from the cone property
of C, one knows that αc1, (1 − α)c2 ∈ C, therefore αc1 + (1 − α)c2 ∈ C + C = C, which
ends the proof.
Definition 2.1.4. Let A ⊂ Rp
be a nonempty set and x ∈ Rp
. One defines the distance
from x to A by the relation:
d(x, A) := inf{kx − ak | a ∈ A}.
We also consider the function dA : Rp
→ R given by
dA(x) := d(x, A).
We now introduce some basic properties of the distance from a point to a
(nonempty) set.
Theorem 2.1.5. Let A ⊂ Rp
, A ≠ ∅. Then:
(i) d (x, A) = 0 if and only if x ∈ cl A.
(ii) The function dA is 1−Lipschitz.
(iii) If A is closed, then for every x ∈ Rp
, there exists ax ∈ A such that d (x, A) =
kx − axk . If, moreover, A is convex, then ax having the previous property is unique and
it is characterized by the relations
(
ax ∈ A
hx − ax, u − axi ≤ 0, ∀u ∈ A.
Proof (i) The following equivalences hold
d (x, A) = 0 ⇔ inf
a∈A
kx − ak = 0
⇔ ∃ (an) ⊂ A with lim
n→∞
kx − ank = 0 ⇔ x ∈ A.
(ii) For every x, y ∈ Rp
and every a ∈ A, these relations hold:
d (x, A) ≤ kx − ak ≤ kx − yk + ky − ak .
As a is taken arbitrary from A, one deduces
d (x, A) ≤ kx − yk + d (y, A) ,
i.e.,
d (x, A) − d (y, A) ≤ kx − yk .
By reversing the roles of x and y, one has:
|d (x, A) − d (y, A)| ≤ kx − yk ,
34. 20 | Nonlinear Analysis Fundamentals
which is the desired conclusion.
(iii) If x ∈ A, then ax := x is the unique element having the previously introduced
property. Let x ∉ A. As d(x, A) is a real number, there exists r 0 such that A1 :=
A ∩ D (x, r) ≠ ∅. Since A1 is a compact set and the function g : A1 → R, g (y) =
d (x, y) is continuous, according to Weierstrass Theorem, g attains its minimum on
A1, i.e., there exists ax ∈ A1 with g (ax) = infy∈A1
g (y) = d(x, A1). Now, one can check
that d (x, A1) = d (x, A) and the first conclusion follows. Suppose that, moreover, A
is convex. If x ∈ A, there is nothing to prove. Take x ∉ A. Consider a1, a2 ∈ A with
d (x, A) = kx − a1k = kx − a2k . Using the parallelogram law we know:
(x − a1) + (x − a2)
2
+ (x − a1) − (x − a2)
2
= 2 kx − a1k2
+ 2 kx − a2k2
,
i.e.,
k2x − a1 − a2k2
+ ka2 − a1k2
= 4d2
(x, A),
and dividing by 4 one gets
x −
a1 + a2
2
2
+ 4−1
ka2 − a1k2
= d2
(x, A).
Since A is convex, 2−1
(a1+a2) ∈ A, hence x −
a1 + a2
2
2
≥ d2
(x, A). This relation and
the previous equality show that ka2 − a1k = 0, hence a1 = a2. The proof of uniqueness
is now complete. Let us prove now that ax verifies the relation hx − ax, u − axi ≤ 0 for
any u ∈ A. For this, take u ∈ A. Then for every α ∈ (0, 1], one has
v = αu + (1 − α)ax ∈ A.
Hence,
kx − axk ≤ x − αu − (1 − α)ax = x − ax − α(u − ax) ,
and, consequently,
kx − axk2
≤ kx − axk2
− 2α hx − ax, u − axi + α2
ku − axk2
.
After reducing terms and dividing by α 0, we can see that
0 ≤ −2 hx − ax, u − axi + α ku − axk2
.
If we let α → 0, the desired inequality follows. For the converse, if an element a ∈ A
satisfies hx − a, u − ai ≤ 0 for any u ∈ A, then for every v ∈ A one has
kx − ak2
− kx − vk2
= 2 hx − a, v − ai − ka − vk2
≤ 0,
hence a coincides with ax. The proof is now complete.
In the case when A is closed, then for x ∈ Rp
one denotes the projection set of x
on A by
prA x :=
a ∈ A | d(x, A) = kx − ak .
35. Convex Sets and Cones | 21
If, moreover, A is convex, then, according to the above theorem, this set consists of
only one element, which we still denote by prA x, and we call this the projection of x
on A.
Let S ⊂ Rp
be a nonempty set. The polar of S is the set
S−
:= {u ∈ Rp
| hu, xi ≤ 0, ∀x ∈ S}.
It is easy to observe that S−
is a closed convex cone and that, in general, S ⊂ (S−
)−
. If
we consider the reverse inclusion, the next result follows.
Theorem 2.1.6. Let C ⊂ Rp
be a closed convex cone. Then C = (C−
)−
.
Proof Consider z ∈ (C−
)−
and z = prC z. We will prove that z = z. From the last part of
Theorem 2.1.5, for any c ∈ C, one has
hz − z, c − zi ≤ 0.
As 0 ∈ C and 2z ∈ C, we deduce
− hz, z − zi ≤ 0, hz, z − zi ≤ 0
hence
hz − z, ci ≤ 0
for every c ∈ C, i.e., z − z ∈ C−
. As z ∈ (C−
)−
, one gets
hz, z − zi ≤ 0.
But
kz − zk2
= hz − z, zi − hz − z, zi ≤ 0
which means that z = z, so z ∈ C. This establishes the theorem.
Example 2.1.7. 1. Consider S = {(x, 0) ∈ R2
| x ≥ 0}. One can observe that S−
=
{(x, y) ∈ R2
| x ≤ 0}. Obviously, (S−
)−
= S−
.
2. The polar of R2
+ := {(x, y) ∈ R2
| x, y ≥ 0} is R2
− := {(x, y) ∈ R2
| x, y ≤ 0}. The
polar of the set S = {(x, 0) ∈ R2
| x ≥ 0} ∪ {(0, y) ∈ R2
| y ≥ 0} is also R2
−. From this
example one can see that, in general, S−
1 = S−
2 does not imply S1 = S2.
The next result, which has an algebraic character, it was obtained by the Hungarian
mathematician Julius Farkas in 1902.
Theorem 2.1.8 (Farkas’ Lemma). Let n ∈ N*
, (φi)i∈1,n ⊂ L(Rp
, R) and φ ∈ L(Rp
, R).
Then
∀x ∈ Rp
: [φ1(x) ≤ 0, . . . , φn(x) ≤ 0] ⇒ φ(x) ≤ 0 (2.1.1)
if and only if there exists (αi)i∈1,n ⊂ [0, ∞) such that φ =
Pn
i=1 αiφi.
36. 22 | Nonlinear Analysis Fundamentals
Proof The converse implication is obvious. We prove the other one by induction for
n ≥ 1. Define the proposition P(n) which says that for every φ, φ1, . . . , φn ∈ L(Rp
, R)
satisfying (2.1.1), there exist (αi)i∈1,n ⊂ [0, ∞) such that φ =
Pn
i=1 αiφi.
Let us prove that P(1) is true. Indeed, let φ, φ1 ∈ L(Rp
, R) such that
φ1(x) ≤ 0 ⇒ φ(x) ≤ 0.
If φ = 0 then, obviously, φ = 0φ1. Suppose φ ≠ 0. Then, by the assumption:
φ1(x) = 0 ⇔ [φ1(x) ≤ 0, φ1(−x) ≤ 0] ⇒ [φ(x) ≤ 0, φ(−x) ≤ 0] ⇔ φ(x) = 0,
hence Ker φ1 ⊂ Ker φ. Since φ ≠ 0, one has φ1 ≠ 0, so there exists x1 ∈ Rp
with
φ1(x) = −1. Also by the assumption, φ(x) ≤ 0. Take x ∈ Rp
arbitrarily. Then it is easy
to verify that
x + φ1(x)x ∈ Ker φ1,
hence
x + φ1(x)x ∈ Ker φ,
i.e.,
φ x + φ1(x)x
= 0,
which proves that
φ(x) = −φ(x)φ1(x).
Notice that x was arbitrarily chosen, so the desired relation is proved for α1 := −φ(x) ≥
0.
Suppose now that P(n) is true for a fixed n ≥ 1 and we will try to prove that P(n+1)
is true.
Take φ, φ1, . . . , φn, φn+1 ∈ L(Rp
, R) such that
∀x ∈ Rp
: [φ1(x) ≤ 0, . . . , φn(x) ≤ 0, φn+1(x) ≤ 0] ⇒ φ(x) ≤ 0. (2.1.2)
If
∀x ∈ Rp
: [φ1(x) ≤ 0, . . . , φn(x) ≤ 0] ⇒ φ(x) ≤ 0, (2.1.3)
then, from P(n), there exist (αi)i∈1,n ⊂ [0, ∞) such that φ =
Pn
i=1 αiφi; take αn+1 := 0
and the conclusion follows.
Suppose relation (2.1.3) is not satisfied. Then there exists x ∈ Rp
such that φ(x) 0
and φi(x) ≤ 0 for any i ∈ 1, n. As (2.1.2) holds, φn+1(x) 0; we may suppose (by
multiplying by the appropriate positive scalar) that φn+1(x) = 1. But
φn+1(x − φn+1(x)x) = 0, ∀x ∈ Rp
,
and from (2.1.2) we deduce
∀x ∈ Rp
: [φ1(x − φn+1(x)x) ≤ 0, . . . , φn(x − φn+1(x)x) ≤ 0] ⇒ φ(x − φn+1(x)x) ≤ 0.
(2.1.4)
37. Convex Sets and Cones | 23
Take φ0
i := φi − φi(x)φn+1 for i ∈ 1, n and φ0
:= φ − φ(x)φn+1, and then relation (2.1.4)
becomes
∀x ∈ X : [φ0
1(x) ≤ 0, . . . , φ0
n(x) ≤ 0] ⇒ φ0
(x) ≤ 0.
As P(n) is true, there exist (αi)i∈1,n ⊂ [0, ∞) such that φ0
=
Pn
i=1 αiφ0
i . We deduce that
φ − φ(x)φn+1 =
n
X
i=1
αi
φi − φi(x)φn+1
,
hence φ =
Pn+1
i=1 αiφi, where αn+1 = φ(x) −
Pn
i=1 αiφi(x) ≥ 0 from the choice of x and
from the fact that αi ≥ 0 for any i ∈ 1, n). The proof is now complete.
Throughout this book we shall use several different concepts of tangent vectors to
a set at a point. We introduce now one of these concepts.
Definition 2.1.9. Let M ⊂ Rp
be a nonempty set and x ∈ cl M. One says that a
vector u ∈ Rp
is tangent in the sense of Bouligand to the set M at x if there exist
(tn) ⊂ (0, ∞), tn → 0 and (un) → u such that for any n ∈ N, one has
x + tnun ∈ M.
It is sufficient that the above inclusion holds for every n ∈ N sufficiently large.
Theorem 2.1.10. The set, denoted by TB(M, x), which contains all the tangent vectors
to the set M at x is a closed cone, which we call the Bouligand tangent cone (or the
contingent cone) to the set M at the point x.
Proof Let us prove first that 0 ∈ TB(M, x). If x ∈ M, then the assertion trivially
follows, because it is sufficient to take (un) constantly equal to 0. If x ∉ M, then
there exists (xn)n∈N ⊂ M such that xn → x, and we consider tn :=
p
kxn − xk and
un :=
p
kxn − xk
−1
(xn − x) for every n ∈ N. Since tn → 0 and un → 0, one obtains
the conclusion.
Consider now u ∈ TB(M, x) and λ 0. According to the definition, there exist
(tn) ⊂ (0, ∞), tn → 0 and (un) → u such that for every n ∈ N,
x + tnun ∈ M.
This is equivalent to
x +
tn
λ
(λun) ∈ M.
As
tn
λ
→ 0 and (λun) → λu, one deduces that λu ∈ TB(M, x), hence TB(M, x) is
a cone. We prove that the closure of TB(M, x) is contained in TB(M, x). Take (un) ⊂
38. 24 | Nonlinear Analysis Fundamentals
TB(M, x) and (un) → u. One must prove that u ∈ TB(M, x). For every n ∈ N, there
exist (tk
n)k ⊂ (0, ∞), tk
n
k→∞
→ 0 and (uk
n)
k→∞
→ un such that for every k ∈ N,
x + tk
nuk
n ∈ M.
By using a diagonalization procedure, for every n ∈ N*
, there exists kn ∈ N such that
the next relations hold:
tkn
n
1
n
ukn
n − un ≤
1
n
.
It is easy to observe that the positive sequence (tkn
n )n converges to 0, and using the
inequality
ukn
n − u ≤ ukn
n − un + kun − uk ,
one can deduce that (ukn
n ) → u. Moreover, for every n ∈ N,
x + tkn
n ukn
n ∈ M,
hence u ∈ TB(M, x) and the proof is complete.
Our first example is given next.
Example 2.1.11. 1. Consider the ball M ⊂ R2
, M := {(x, y) ∈ R2
| (x − 1)2
+ y2
≤ 1}.
Then TB(M, (0, 0)) = {(x, y) ∈ R2
| x ≥ 0}.
2. One can easily observe that if C ⊂ Rp
is a closed cone, then TB(C, 0) = C.
Proposition 2.1.12. If ∅ ≠ M ⊂ Rp
and x ∈ cl M, then TB(M, x) = TB(cl M, x). If
x ∈ int M, then TB(M, x) = Rp
.
Proof The inclusion TB(M, x) ⊂ TB(cl M, x) is obvious by the use of M ⊂ cl M. Take
u ∈ TB(cl M, x). There exist (tn) ⊂ (0, ∞), tn → 0 and (un) → u such that for every
n ∈ N,
x + tnun ∈ cl M.
Using the sequence characterization of the closure, for any fixed n, there exist (vk
n)k ⊂
M such that
vk
n
k
→ x + tnun.
As above, for any fixed n, there exists kn ∈ N such that
vkn
n − (x + tnun) ≤ t2
n.
Then, one can write
vkn
n − x
tn
− u ≤
vkn
n − x
tn
− un + kun − uk ≤ tn + kun − uk ,
39. Convex Sets and Cones | 25
hence u0
n :=
vkn
n − x
tn
n→∞
→ u. But
x + tnu0
n = vkn
n ∈ M,
hence u ∈ TB(M, x). The second part of the conclusion easily follows: if x ∈ int M,
then for every u ∈ Rp
and every (tn) ⊂ (0, ∞), tn → 0, one has x + tnun ∈ M for any
n sufficiently large. This shows, in particular, that u ∈ TB(M, x), and the conclusion
follows.
In general, the Bouligand tangent cone is not convex and the relation TB(M, x) =
Rp
can be satisfied, even if x ∉ int M.
Example 2.1.13. 1. Consider the set M ⊂ R2
, M = {(x, y) | x ≥ 0, y = 0} ∪ {(x, y) | x =
0, y ≥ 0}. Then TB(M, (0, 0)) = M is not a convex set.
2. Let set M represent the plane domain bounded by the curve (the cardioid) which
has the parametric representation
(
x = −2 cos t + cos 2t + 1
y = 2 sin t − sin 2t
, t ∈ [0, 2π].
Then TB(M, (0, 0)) = R2
, but (0, 0) ∉ int M.
Figure 2.1: The cardioid.
40. 26 | Nonlinear Analysis Fundamentals
Proposition 2.1.14. Let A1, A2 ⊂ Rp
be closed sets. Then the next relations hold:
(i) if x ∈ A1 ∩ A2, then TB(A1 ∪ A2, x) = TB(A1, x) ∪ TB(A2, x);
(ii) if x ∈ A1 ∩ A2, then TB(A1 ∩ A2, x) ⊂ TB(A1, x) ∩ TB(A2, x);
(iii) if x ∈ bd A1, then TB(bd A1, x) = TB(A1, x) ∩ TB(Rp
A1, x).
Proof The first two relations easily follow, as well as the inclusion TB(bd A1, x) ⊂
TB(A1, x) ∩ TB(Rp
A1, x) from (iii), which can be proved by using (ii) and Propo-
sition 2.1.12. Let us prove now the other inclusion from (iii). Take u ∈ TB(A1, x) ∩
TB(Rp
A1, x). According to the definition, there exist (tn), (t0
n) ⊂ (0, ∞), tn, t0
n → 0
and (un), (u0
n) → u such that for every n ∈ N,
x + tnun ∈ A1,
and
x + t0
nu0
n ∈ Rp
A1.
If an infinite number of terms from the first or the second relation are on the boundary
of A1, there is nothing to prove. Suppose next, without loss of generality, that for every
n ∈ N, there exists λn ∈ (0, 1) such that
λn (x + tnun) + (1 − λn)(x + t0
nu0
n) ∈ bd A1.
Consider the sequences
(t00
n ) := (λntn + (1 − λn)t0
n) ⊂ (0, ∞)
(u00
n ) :=
tnλn
t00
n
un +
t0
n(1 − λn)
t00
n
u0
n.
It is clear that (t00
n ) → 0. On the other hand,
u00
n − u ≤ kun − uk + u0
n − u ,
hence (u00
n ) → u. Since
x + t00
n u00
n ∈ bd A1,
one gets the desired conclusion.
Denote by NB(M, x) the polar of TB(M, x) (i.e., NB(M, x) := TB(M, x)−
), and we call
this set the Bouligand normal cone to M at x.
If the set M is convex, then the Bouligand tangent and normal cones have a special
form.
Proposition 2.1.15. Let ∅ ≠ M ⊂ Rp
be a convex set and x ∈ M. Then
TB(M, x) = cl R+(M − x),
and
NB(M, x) = {u ∈ Rp
| hu, c − xi ≤ 0, ∀c ∈ M}.
41. Convex Sets and Cones | 27
Proof Take c ∈ M and d := c − x. Consider (tk)k → 0. Then
x + tkd = (1 − tk)x + tkc ∈ M,
hence M − x ⊂ TB(M, x). Since TB(M, x) is a closed cone, one gets that cl R+(M − x) ⊂
TB(M, x). Take u ∈ TB(M, x). Then there exist (tk) ⊂ (0, ∞), tk → 0 and (uk) → u
such that for every k ∈ N,
xk: = x + tkuk ∈ M.
Hence u = limk
xk − x
tk
. But
xk − x
tk
k
⊂ R+(M − x). One can deduce that TB(M, x) ⊂
cl R+(M − x). Recall that, by definition,
NB(M, x) = TB(M, x)−
= {u ∈ Rp
| hu, vi ≤ 0, ∀v ∈ TB(M, x)}.
Now, taking into account the particular form of TB(M, x), the conclusion follows.
For a reason we make clear later on, in the case of convex sets, we do not use the
subscript B in the notation of these cones.
Example 2.1.16. We want to compute the tangent and normal cones, at different points,
to the set M ⊂ Rp
,
M =
(
x = (x1, x2, ..., xp) ∈ Rp
| xi ≥ 0, ∀i ∈ 1, p,
p
X
i=1
xi = 1
)
,
which is called the unit simplex. This set is convex and closed. According to the previous
result, for every x ∈ M,
T(M, x) = cl R+(M − x)
= cl
u ∈ Rp
| ∃α ≥ 0, x ∈ M, u = α(x − x) .
Take u from the right-hand set. It is clear that, on one hand,
Pp
i=1 ui = 0, and, on the
other hand, if xi = 0, then ui ≥ 0. Denote by I(x) :=
i ∈ 1, p | xi = 0 . It follows that
T(M, x) ⊂
(
u ∈ Rp
|
p
X
i=1
ui = 0 and ui ≥ 0, ∀i ∈ I(x)
)
.
Let us now prove the reverse inclusion. It is easy to verify that the right-hand set is closed.
Take u from this set. If u = 0, then, obviously, u ∈ T(M, x). If u ≠ 0, then we must prove
that there exists α 0 such that x+αu ∈ M. On one hand, it is clear that
Pp
i=1(xi +αui) =
1 is satisfied for any α. If there is no i with ui 0, then it is also easy to observe that
xi + αui ≥ 0, for any i ∈ 1, p, hence u ∈ T(M, x). Suppose that the set J of indices for
which uj 0 is nonempty. Then J ⊂ 1, p I(x), hence xj 0 for any j ∈ J. One can
choose the positive α such that
α min{−u−1
j xj | j ∈ J},
42. 28 | Nonlinear Analysis Fundamentals
and again one has xi + αui ≥ 0, for any i ∈ 1, p. Hence, u ∈ T(M, x), and the double
inclusion follows.
We prove next that
N(M, x) =
(a, a, ..., a) ∈ Rp
| a ∈ R +
v ∈ Rp
| vi ≤ 0, ∀i ∈ I(x), vi = 0, i ∉ I(x) .
For this, consider the elements
a0 = (1, 1, ..., 1), a1 = −(1, 0, ..., 0), ..., an = −(0, 0, ..., 1)
and observe that T(M, x) can be equivalently written as:
T(M, x) =
u ∈ Rp
| ha0, ui ≤ 0, h−a0, ui ≤ 0, hai, ui ≤ 0, ∀i ∈ I(x) .
The polar of this set is
N(M, x) =
αa0 − βa0 +
X
i∈I(x)
αiai | α, β, αi ≥ 0, ∀i ∈ I(x)
.
Indeed, the fact that the right-hand set is contained in the normal cone is obvious, and
the reverse inclusion follows from Farkas’ Lemma (Theorem 2.1.8). We now obtain the
desired form of the normal cone.
At the end of this section, we discuss the concepts of convex hull and conic hull of a
set. Let A ⊂ Rp
be a nonempty set. The convex hull of A is the set
conv A =
( n
X
i=1
αixi | n ∈ N*
, (αi)i∈1,n ⊂ [0, ∞),
n
X
i=1
αi = 1, (xi)i∈1,n ⊂ A
)
.
It is not difficult to see that conv A is a convex set which contains A. One can easily
verify that conv A is the smallest set (in the sense of inclusion) with these properties
(see Problem 7.26).
The conic hull of the set A is
cone A := [0, ∞)A := {αx | α ≥ 0, x ∈ A} .
In fact, cone A is the smallest cone which contains A.
We give next two results concerning these sets. The first one refers to the structure
of the set conv A and it is called the Carathéodory Theorem, after the name of the Greek
mathematician Constantin Carathéodory, who proved this result in 1911 for compact
sets.
Theorem 2.1.17 (Carathéodory Theorem). Let A ⊂ Rp
be a nonempty set. Then
conv A =
(p+1
X
i=1
αixi | (αi)i∈1,p+1 ⊂ [0, ∞),
p+1
X
i=1
αi = 1, (xi)i∈1,p+1 ⊂ A
)
.
43. Convex Sets and Cones | 29
Proof We must prove that every element from conv A can be written as a combination
of at most p + 1 elements from A. Consider x ∈ conv A. According to the definition
of conv A, x can be written as a convex combination of elements from A. Suppose,
by means of contradiction, that the minimal number of elements from A which can
form a convex combination equal to x is n p + 1. So there exist x1, x2, ..., xn ∈ A,
α1, α2, ..., αn ∈ (0, 1) with
Pn
i=1 αi = 1 such that
Pn
i=1 αixi = x. Then the elements
(xi − xn)i=1,n−1 are linearly dependent (their number is greater than the dimension p
of the space), so there exist (λi)i=1,n−1, not all equal to 0, such that
n−1
X
i=1
λi(xi − xn) = 0,
which means
n−1
X
i=1
λixi −
n−1
X
i=1
λi
!
xn = 0.
By denoting −
Pn−1
i=1 λi
= λn, one has
Pn
i=1 λi = 0 and
Pn
i=1 λixi = 0. Then for every
t ∈ R,
x =
n
X
i=1
αixi + t
n
X
i=1
λixi =
n
X
i=1
(αi + tλi)xi
and
n
X
i=1
(αi + tλi) = 1.
As
Pn
i=1 λi = 0 and there is at least one nonzero element, there exists at least one
negative value among the numbers (λi)i∈1,n. Denote t := min{−αiλ−1
i | λi 0}. Then
all the values (αi + tλi) are in the interval [0, ∞), and the corresponding value of the
index which gives the minimum from above is zero, whence x is a convex combina-
tion of less than n elements from A, contradicting the minimality of n. Therefore, the
assumption that we made was false, and the conclusion follows.
We now discuss the necessary conditions one needs in order that the conic hull of
a set is closed. This does not happens automatically, as one can see from the example
given by A ⊂ R2
, A := {(x, y) ∈ R2
| (x − 1)2
+ y2
= 1}, for which cone A = {(x, y) ∈
R2
| x 0} ∪ {(0, 0)}.
First, one defines for a nonempty set A ⊂ X the asymptotic cone of A as
A∞
= {u ∈ X | ∃(tn) → 0, ∃(an) ⊂ A, tnan → u}.
It is clear, by repeating the arguments from the case of the Bouligand tangent cone,
that A∞
is a closed cone. If A is bounded, then A∞
= {0}, and the converse also holds
(if A would contain an unbounded sequence (an), then
an
kank
would also have a
44. 30 | Nonlinear Analysis Fundamentals
subsequence which converges to a nonzero element, which must be from A∞
). Let us
observe also that if A is a cone, then A∞
= TB(A, 0) = cl A.
The next result concerns decomposition.
Theorem 2.1.18. Let A ⊂ Rp
be a nonempty closed set.
(i) If 0 ∉ A, then cl cone A = cone A ∪ A∞
.
(ii) If 0 ∈ A, then cl cone A = cone A ∪ A∞
∪ TB(A, 0).
Proof (i) Suppose that 0 ∉ A. It is clear that cone A ⊂ cl cone A. From the def-
inition of A∞
, one also has that A∞
⊂ cl cone A. For the reverse inclusion, take
(un) ⊂ cone A, un → u. We must prove that u ∈ cone A ∪ A∞
. If u = 0, then the
relation u ∈ cone A is obvious. Suppose that u ≠ 0, then for every n ∈ N, there ex-
ist tn ≥ 0 and an ∈ A such that un = tnan. If (an) is unbounded, one can pass to a
subsequence (ank ), where kank k → ∞. Hence tnk → 0, and u ∈ A∞
. Suppose (an) is
bounded. Since 0 ∉ A and A is closed, there exists γ 0 such that kank ≥ γ for every
n. One can deduce that (tn) is bounded, so it converges (on a subsequence (tnk ), even-
tually) to a number t ≥ 0. If t = 0, then unk → 0 = u (a situation which is excluded at
this point of the proof). Accordingly, t 0 and
ank − t−1
u = t−1
ktank − uk = t−1
tnk ank − u + (t − tnk )ank
≤ t−1
ktnk ank − uk + |t − tnk | t−1
kank k → 0.
Hence ank → t−1
u and since A is closed, u ∈ cone A. The proof of this part is complete.
(ii)Theinclusion cone A∪A∞
∪TB(A, 0) ⊂ cl cone A isobvious.Take u ∈ cl cone A.
In the above it is possible as well that (an) converges to 0. Then tn → ∞, hence u ∈
TB(A, 0), which completes the proof.
We finish with the following characterization result:
Corollary 2.1.19. Let A ⊂ Rp
be a nonempty closed set.
(i) If 0 ∉ A, then cone A is closed if and only if A∞
⊂ cone A.
(ii) If 0 ∈ A, then cone A is closed if and only if A∞
∪ TB(A, 0) ⊂ cone A.
2.2 Convex Functions
2.2.1 General Results
In this section we present the special class of convex functions. These functions are
defined on convex sets.
Definition 2.2.1. Let D ⊂ Rp
be a convex set. One says that a function f : D → R is
convex if
f(λx + (1 − λ)y) ≤ λf(x) + (1 − λ)f(y), ∀x, y ∈ D, ∀λ ∈ [0, 1]. (2.2.1)
45. Convex Functions | 31
It is clear that in the above definition it is sufficient to take λ ∈ (0, 1).
As said before, in R the convex sets are exactly the intervals. In this framework, the
convexity has the following geometric meaning: for every two points x, y ∈ D, x y,
the graph of the restriction of f to the [x, y] interval lies below the line segment joining
the points (x, f(x)) and (y, f(y)). This can be written as follows: for every u ∈ [x, y],
f(u) ≤ f(x) +
f(y) − f(x)
y − x
(u − x), (2.2.2)
inequality which can be deduced from (2.2.1) by replacing λ with the value given by
the relation u = λx +(1− λ)y. Therefore, (2.2.1) and (2.2.2) are equivalent (for functions
defined on R).
Definition 2.2.2. Let D ⊂ Rp
be a convex set. One says that a function f : D → R is
concave if −f is convex.
All of the properties of concave functions can be easily deduced from the similar prop-
erties of the convex functions, so in what follows we will consider only the later case.
We first deduce some general properties of the convex functions.
Proposition 2.2.3. Let D ⊂ Rp
be a convex set and f : D → R. The following relations
are equivalent:
(i) f is convex;
(ii) the epigraph of f ,
epi f := {(x, t) ∈ D × R | f(x) ≤ t},
is a convex subset of Rp
× R;
(iii) for any x, y ∈ D, define
Ix,y := {t ∈ R | tx + (1 − t)y ∈ D};
then the function φx,y : Ix,y→ R, φx,y(t) = f(tx + (1 − t)y) is convex.
Proof We prove first the implication from (i) to (ii). Take λ ∈ (0, 1) and (x, t), (y, s) ∈
epi f . By the convexity of D, one knows that λx +(1− λ)y ∈ D, and using the convexity
of f , one can say:
f(λx + (1 − λ)y) ≤ λf(x) + (1 − λ)f(y) ≤ λt + (1 − λ)s,
i.e., λx + (1 − λ)y, λt + (1 − λ)s
∈ epi f . Therefore, epi f is a convex set.
We prove now the converse implication. Take x, y ∈ D and λ ∈ [0, 1]. Then
(x, f(x)), (y, f(y)) ∈ epi f and by assumption, λ(x, f(x)) + (1 − λ)(y, f(y)) ∈ epi f , hence
f(λx + (1 − λ)y) ≤ λf(x) + (1 − λ)f(y),
which shows that f is a convex function.
46. 32 | Nonlinear Analysis Fundamentals
We prove now the equivalence between (i) and (iii). Observe first that Ix,y is an
interval which contains [0, 1]. Suppose that f is convex and take u, v ∈ Ix,y, λ ∈ [0, 1].
One knows that:
φx,y(λu + (1 − λ)v) = f([λu + (1 − λ)v]x + [1 − λu − (1 − λ)v]y)
= f(λ(ux + (1 − u)y) + (1 − λ)(vx + (1 − v)y))
≤ λf(ux + (1 − u)y) + (1 − λ)f(vx + (1 − v)y)
= λφx,y(u) + (1 − λ)φx,y(v).
For the converse implication, take x, y ∈ D and t ∈ [0, 1]. Then φx,y is convex, hence
for every λ ∈ [0, 1], u, v ∈ Ix,y
φx,y(λu + (1 − λ)v) ≤ λφx,y(u) + (1 − λ)φx,y(v)
= λf(ux + (1 − u)y) + (1 − λ)f(vx + (1 − v)y).
By taking u = 1, v = 0, λ = t we deduce that
φx,y(t) ≤ tf(x) + (1 − t)f(y),
hence f is convex.
Theorem 2.2.4. Let D ⊂ Rp
be a convex set and f : D → R be a convex function. Then
f is continuous at every interior point of D.
Proof Take x ∈ int D. A translation permits us to consider the case x = 0. We prove
first that f is bounded on a neighborhood of 0. If we denote by (ei)i∈1,p the canonical
base of Rp
, then there exists an a 0 such that aei and −aei are in D for any i ∈ 1, p.
Under these conditions, the set
V :=
(
x ∈ Rp
| x =
p
X
i=1
xiei, |xi|
a
p
, ∀i ∈ 1, p
)
is a neighborhood of 0 contained in D. For x ∈ V, there exist (xi)i∈1,p with |xi| a
p ,
i ∈ 1, p and x =
Pp
i=1 xiei. Suppose first that xi ≠ 0 for any i ∈ 1, p. One has
f(x) = f
p
X
i=1
xiei
!
= f
p
X
i=1
|xi|
a
a
xi
|xi|
ei + 1 −
p
X
i=1
|xi|
a
!
0
!
≤
p
X
i=1
|xi|
a
f
a
xi
|xi|
ei
+ 1 −
p
X
i=1
|xi|
a
!
f(0)
≤ max{f (aei) , f(−aei) | i ∈ 1, p} + f(0) .
We can now observe that if there are indices i for which xi = 0, then these can be
excluded from the above calculations, and the estimation holds.
47. Convex Functions | 33
Since the right-hand part is a constant (which we denote by M), the proof is fin-
ished. Take ε ∈ (0, 1) and U a symmetric neighborhood of 0 such that ε−1
U ⊂ V.
Then, for any x ∈ U,
f(x) = f
ε(ε−1
x) + (1 − ε)0
≤ εf(ε−1
x) + (1 − ε)f(0) ≤ εM + (1 − ε)f(0),
i.e.,
f(x) − f(0) ≤ εM − εf(0).
From the fact that U is symmetric, one deduces that for every x ∈ U
f(−x) ≤ εM + (1 − ε)f(0).
Moreover,
f(0) = f
1
2
x +
1
2
(−x)
≤
1
2
f(x) +
1
2
f(−x) ≤
1
2
f(x) +
1
2
εM + (1 − ε)f(0)
,
hence
f(0) − f(x) ≤ εM − εf(0).
This relation can be combined with the similar one from above, and gives
f(x) − f(0) ≤ εM − εf(0).
This inequality proves the continuity of f at 0.
We want to emphasize now some characterizations of differentiable convex func-
tions. Some preliminary results on convex functions defined on a real intervals are
necessary.
Proposition 2.2.5. Let I ⊂ R be an interval and f : I → R be a function. The next
relations are equivalent:
(i) f is convex;
(ii) for every x1, x2, x3 ∈ I satisfying the relation x1 x2 x3 one has
f(x2) − f(x1)
x2 − x1
≤
f(x3) − f(x1)
x3 − x1
≤
f(x3) − f(x2)
x3 − x2
;
(iii) for every a ∈ int I, the function g : I {a} → R given by
g(x) =
f(x) − f(a)
x − a
is increasing.
Proof We prove the (i) ⇒ (ii) implication. Take λ =
x2 − x1
x3 − x1
∈ (0, 1). Then the equality
x2 = λx3 + (1 − λ)x1 holds and one must prove now that
f(x2) − f(x1)
λ(x3 − x1)
≤
f(x3) − f(x1)
x3 − x1
≤
f(x3) − f(x2)
(1 − λ)(x3 − x1)
.
48. 34 | Nonlinear Analysis Fundamentals
After some calculations, one can show that:
f(x2) ≤ λf(x3) + (1 − λ)f(x1).
The proof of the implication (ii) ⇒ (i) follows the inverse path of the proof of (i) ⇒ (ii),
hence (i) and (ii) are equivalent.
In order to prove (ii) ⇒ (iii), we fix x1, x2 ∈ I {a} with x1 x2 and we find three
situations. If x1 x2 a, then we apply (ii) for the triplet (x1, x2, a). If x1 a x2,
then we apply (ii) for the triplet (x1, a, x2). Finally, if a x1 x2, then we apply (ii)
for the triplet (a, x1, x2).
We prove now (iii) ⇒ (i). Take x, y ∈ I with x y and λ ∈ (0, 1). Then x
λx + (1 − λ)y y, and by applying (iii) with a = λx + (1 − λ)y, one deduces
f(x) − f(λx + (1 − λ)y)
x − λx − (1 − λ)y
≤
f(y) − f(λx + (1 − λ)y)
y − λx − (1 − λ)y
.
After some calculations, the relation follows from the definition of convexity. The fact
that this relation holds for any x, y ∈ I with x y and for any λ ∈ (0, 1) is sufficient to
prove the desired assertion. The proof is complete.
Proposition 2.2.6. Let I ⊂ R be an interval and f : I → R be a convex function. Then f
admits lateral derivatives in every interior point of I and for every x, y ∈ int I with x y,
one has
f0
−(x) ≤ f0
+(x) ≤ f0
−(y) ≤ f0
+(y).
Proof Fix a ∈ int I. Since the function g : I {a} → R given by
g(x) =
f(x) − f(a)
x − a
is increasing (see the previous result) one deduces that g admits finite lateral limits,
which implies the existence of the lateral derivatives of f at a. Moreover, f0
−(a) ≤ f0
+(a).
For x, y ∈ int I, x y and for any u, v ∈ (x, y), u ≤ v, by using again the argument
given by the last conclusion of Proposition 2.2.5, one deduces that
f(u) − f(x)
u − x
≤
f(v) − f(x)
v − x
=
f(x) − f(v)
x − v
≤
f(y) − f(v)
y − v
=
f(v) − f(y)
v − y
.
Passing to the limit for u → x and v → y, one gets f0
+(x) ≤ f0
−(y).
Here we characterize differentiable convex functions of one variable.
Theorem 2.2.7. Let I be an open interval and f : I → R be a function.
(i) If f is differentiable on I, then f is convex if and only if f0
is increasing on I.
(ii) If f is twice differentiable on I, then f is convex if and only if f00
(x) ≥ 0 for every
x ∈ I.
49. Convex Functions | 35
Proof In the case of real functions of one variable, the equivalence between the mono-
tonicity of f0
and the sign of f00
is sufficient to prove that f is convex if and only if f0
is
increasing on I. If f is convex, the monotonicity of the derivative follows from Propo-
sition 2.2.6. Conversely, suppose that f0
is increasing and we prove that f is convex.
Take a, b ∈ I. Define g : [a, b] → R given by
g(x) = f(x) − f(a) − (x − a)
f(b) − f(a)
b − a
.
Obviously, g(a) = g(b) = 0, and
g0
(x) = f0
(x) −
f(b) − f(a)
b − a
.
The function f satisfies the conditions of Lagrange Theorem on [a, b], hence there
exists c ∈ (a, b) such that
f(b) − f(a)
b − a
= f0
(c).
Consequently, g0
(x) = f0
(x) − f0
(c). From the monotonicity of f0
, we deduce that g is
decreasing on (a, c) and increasing on (c, b), and since g(a) = g(b) = 0, we know that
g is negative on the whole interval [a, b]. Take x ∈ (a, b). Then there exists λ ∈ (0, 1)
such that
x = λa + (1 − λ)b.
By replacing x in the expression of g and taking into account that g(x) ≤ 0, we deduce
that
f(λa + (1 − λ)b) − f(a) − (1 − λ)(b − a)
f(b) − f(a)
b − a
≤ 0,
relation which reduces to the definition of the convexity.
Example 2.2.8. Based on the above result, one deduces the convexity of the following
functions: f : R → R, f(x) = ax + b, with a, b ∈ R; f : (0, ∞) → R, f(x) = − ln x;
f : (0, ∞) → R, f(x) = x ln x; f : (0, ∞) → R, f(x) = xa
, a ≥ 1; f : R → R, f(x) = ex
;
f : ( − 1, 1) → R, f(x) = −
√
1 − x2; f : (0, π) → R, f(x) = sin−1
x.
Another example is given by the next result.
Proposition 2.2.9. Let D ⊂ Rp
be a nonempty convex set. Then the function dD : Rp
→
R given by dD(x) = d(x, D) is convex.
Proof Take x, y ∈ Rp
and α ∈ [0, 1]. For any ε 0, there exist dx,ε, dy,ε ∈ D such that
kdx,ε − xk dD(x) + ε
kdy,ε − yk dD(y) + ε.
50. 36 | Nonlinear Analysis Fundamentals
Using the convexity of D, one knows:
dD(αx + (1 − α)y) ≤ αx + (1 − α)y − (αdx,ε + (1 − α)dy,ε)
≤ α kdx,ε − xk + (1 − α) kdy,ε − yk
αdD(x) + (1 − α)dD(y) + ε.
As ε is arbitrarily chosen, we may pass to the limit for ε → 0 and the conclusion
follows.
We now characterize differentiable convex functions in the general case.
Theorem 2.2.10. Let D ⊂ Rp
be an open convex set and f : D → R be a function.
(i) If f is differentiable on D, then f is convex if and only if for any x, y ∈ D,
f(y) ≥ f(x) + ∇f(x)(y − x).
(ii) If f is twice differentiable on D, then f is convex if and only if for every x ∈ D
and y ∈ Rp
, one has
∇2
f(x)(y, y) ≥ 0.
Proof (i) Consider first the case when p = 1. Fix x ∈ D with y ≠ x and take λ ∈ (0, 1].
Since f is convex, we get
f (x + λ (y − x)) = f ((1 − λ) x + λy)
≤ (1 − λ) f(x) + λf(y) = f(x) + λ (f (y) − f (x)) .
Consequently,
f (x + λ (y − x)) − f (x)
λ (y − x)
· (y − x) ≤ f (y) − f (x) .
Passing to the limit for λ → 0, one gets f0
(x) (y − x) ≤ f (y) − f (x).
We pass now to the general case. For x ∈ D, consider the function φy,x from Propo-
sition 2.2.3, which we already know that is convex. Moreover, φy,x is differentiable on
the open interval Iy,x, and
φ0
y,x(t) = ∇f(ty + (1 − t)x)(y − x).
According to the preceding step,
φy,x(1) ≥ φy,x(0) + φ0
y,x(0)
which means that
f(y) ≥ f(x) + ∇f(x)(y − x).
51. Convex Functions | 37
Conversely, fix x, y ∈ D and λ ∈ [0, 1]. Therefore, by assumption,
f (x) ≥ f λx + (1 − λ)y
+ (1 − λ)∇f λx + (1 − λ)y
(x − y)
and
f (y) ≥ f λx + (1 − λ)y
+ λ∇f λx + (1 − λ)y
(y − x) .
Multiplying the first inequality by λ, the second one by (1 − λ), and summing up
the new inequalities, one obtains
λf (x) + (1 − λ)f (y) ≥ f λx + (1 − λ)y
,
which proves that f is convex.
(ii) The case p = 1 is proved in Theorem 2.2.7. Now, in order to pass to the general
case, take x ∈ D, y ∈ Rp
. Suppose f is convex. Since D is open, there exists an α 0
such that u := x + αy ∈ D. According to the assumption, φu,x is convex, and taking
into account the case that we have already studied, φ00
u,x(t) ≥ 0 for any t ∈ Iu,x. For
t = 0, one deduces that
0 ≤ φ00
u,x(0) = ∇2
f(x)(u − x, u − x),
and the conclusion follows. Conversely, for x, y ∈ D and t ∈ Ix,y, φ00
x,y(t) ≥ 0. From the
case p = 1, we get that φx,y is convex, hence f is convex. The proof is now complete.
From these results, one may observe that some properties of the convex functions
have a global character, an aspect which will persist in subsequent sections.
At the end of this subsection, we will discuss a property which is stronger than
convexity.
Definition 2.2.11. Let D ⊂ Rp
be a convex set. One says that a function f : D → R is
strictly convex if
f(λx + (1 − λ)y) λf(x) + (1 − λ)f(y), ∀x, y ∈ D, x ≠ y, ∀λ ∈ (0, 1).
Definition 2.2.12. Let D ⊂ Rp
be a convex set. One says that a function f : D → R is
strictly concave if −f is strictly convex.
Every strictly convex function is convex, but the converse is false. To see this, consider
a convex function which is constant on an interval. Again, the properties of the strictly
concave functions easily follow from the corresponding ones of the strictly convex
functions.
By the use of very similar arguments as in the proofs of preceding results, one can
deduce the next characterizations.
52. 38 | Nonlinear Analysis Fundamentals
Theorem 2.2.13. Let I be an open interval and f : I → R be a differentiable function.
The next assertions are equivalent:
(i) f is strictly convex;
(ii) f (x) f (a) + f0
(a) (x − a) , for any x, a ∈ I, x ≠ a;
(iii) f0
is strictly increasing.
If, moreover, f is twice differentiable (on I), then one more equivalence holds :
(iv) f00
(x) ≥ 0 for any t ∈ I and {x ∈ I | f00
(x) = 0} does not contain any proper
interval.
Example 2.2.14. Using this result, one gets the strict convexity of the following func-
tions: f : (0, ∞) → R, f(x) = − ln x; f : (0, ∞) → R, f(x) = x ln x; f : (0, ∞) → R,
f(x) = xa
, a 1; f : R → R, f(x) = ex
; f : (0, ∞) → R, f(x) = (1 + xp
)
1
p , p 1.
Theorem 2.2.15. Let D ⊂ Rp
an open convex set and f : D → R be a differentiable
function. The next assertions are equivalent:
(i) f is strictly convex;
(ii) f (x) f (a) + ∇f(a) (x − a) , for any x, a ∈ D, x ≠ a.
If, moreover, f is twice differentiable (on D), then the preceding two items are implied
by the relation:
(iii) ∇2
f(x)(y, y) 0 for any x ∈ D and y ∈ Rp
{0}.
2.2.2 Convex Functions of One Variable
In this subsection we will focus on some properties and applications of convex func-
tions defined on real intervals, even if some of the results hold in more general situa-
tions.
The class of convex functions is stable under several algebraic operations, an as-
pect which makes it very useful. Here are some of these operations:
Proposition 2.2.16. Let I, J ⊂ R be intervals.
(i) Let n ∈ N*
and f1, f2, ..., fn : I → R be convex functions, and λ1, λ2, ..., λn ≥
0. Then
Pn
i=1 λifi is convex. If at least one of the functions is strictly convex, and the
corresponding scalar is not zero, then
Pn
i=1 λifi is strictly convex.
(ii) Let f : I → J be (strictly) convex, g : J → R be convex and (strictly) increasing.
Then g ◦ f is (strictly) convex.
(iii) Let f : I → J be strictly decreasing, (strictly) convex, and surjective. Then f−1
is
(strictly) convex.
The next result emphasizes some monotonicity properties of convex functions of one
variable.
53. Convex Functions | 39
Theorem 2.2.17. Let I be a nondegenerate interval (i.e., not a singleton set) and f : I →
R be convex. Then either f is monotone on int I, or there exists x ∈ int I such that f is
decreasing on I ∩ (−∞, x], and increasing on I ∩ [x, ∞).
Proof Because of relation (2.2.2), it is sufficient to restrict our attention to the case
when I is open, i.e., I = int I. Suppose that f is not monotone on I. Then there exist
a, b, c ∈ I, a b c such that f(a) f(b) f(c) or f(a) f(b) f(c). The second
situation cannot hold, since in that case, using (2.2.2), one would have
f(b) ≤ f(a) +
f(c) − f(a)
c − a
(b − a) =
f(a)(c − b) + f(c)(b − a)
c − a
f(b).
Hence, f(a) f(b) f(c). As f is continuous on [a, c], its minimum on this interval
must be attained at a point x (from the Weierstrass Theorem). Take x ∈ I ∩ (−∞, a).
According to Proposition 2.2.5,
f(x) − f(x)
x − x
≤
f(a) − f(x)
a − x
,
i.e.,
(x − a)f(x) ≥ (x − a)f(x) + (x − x)f(a) ≥ (x − a)f(x),
hence f(x) ≤ f(x). Similarly, one can prove that f(x) ≤ f(x) for x ∈ I ∩ (c, ∞). It follows
that f(x) = inf f(I). We next prove that f is decreasing on I ∩ (−∞, x). Take u, v ∈
I ∩ (−∞, x), u v. On one hand,
f(v) − f(u)
v − u
≤
f(x) − f(u)
x − u
,
and on the other hand,
f(x) − f(u)
x − u
=
f(u) − f(x)
u − x
≤
f(v) − f(x)
v − x
≤ 0,
hence f(v) − f(u) ≤ 0, which is exactly what we wanted to prove. Similarly, one can
show that f is increasing on I ∩ (x, +∞). The continuity of f on I finalizes the proof.
As we saw before, a convex function defined on an interval can have discontinu-
ities only at the extremities of the interval. The previous theorem allows us to consider
a different function in those eventual discontinuity points, without losing the convex-
ity. In this way one obtains the next consequence.
Corollary 2.2.18. Let a, b ∈ R, a b and f : [a, b] → R be a convex function. Then
there exist limx→a+ f(x) and limx→b− f(x), and the function
f(x) =
limx→a+ f(x), x = a
f(x), x ∈ (a, b)
limx→b− f(x), x = b
is convex and continuous on [a, b].
54. 40 | Nonlinear Analysis Fundamentals
From the proof of Theorem 2.2.17 one can also obtain the next result, which will be
restated in a more general framework in the next chapter.
Corollary 2.2.19. Let I ⊂ R be a nondegenerate interval. If f : I → R is a convex and
non-monotone function, then it has a global minimum on int I.
The next proposition, sometimes called the Jensen inequality, follows by applying the
definitions and the mathematical induction principle.
Proposition 2.2.20. Let D ⊂ Rp
be a convex set and f : D → R. If the function f is
convex, then
f(λ1x1 + ... + λmxm) ≤ λ1f(x1) + ... + λmf(xm)
for any m ∈ N*
, x1, ..., xm ∈ D, λ1, ..., λm ≥ 0, λ1 + ... + λm = 1. The inequality is
strict if the function f is strictly convex, at least two of the points (xk) are different and
the corresponding scalars (λk) are strictly positive.
Actually, for the convexity of a continuous function it is sufficient that the inequality
from the definition is satisfied for λ = 2−1
.
Theorem 2.2.21. Let I ⊂ R be an interval and f : I → R be a continuous function. The
function f is convex if and only if
f
x + y
2
≤
f(x) + f(y)
2
, ∀x, y ∈ I. (2.2.3)
Proof The necessity of the condition (2.2.3) is obvious. Let us prove that it is also suf-
ficient. Suppose, by contradiction, that f is not convex, which means (2.2.2) is not
satisfied. Then there exist x, y ∈ I, x y and u ∈ (x, y) such that
f(u) f(x) +
f(y) − f(x)
y − x
(u − x). (2.2.4)
Observe by the relation (2.2.4) that the cases u = x and u = y cannot hold. Consider
then g : [x, y] → R,
g(t) = f(t) − f(x) −
f(y) − f(x)
y − x
(t − x),
which is continuous and satisfies the relations g(x) = g(y) = 0. By (2.2.4) and Weier-
strass’ Theorem, there exists z ∈ (x, y) such that g(z) = supt∈[x,y] g(t) 0. Denote
by
w := inf{z ∈ (x, y) | g(z) = sup
t∈[x,y]
g(t)}.
By the continuity of g, it follows that g(w) = supt∈[a,b] g(t) 0, and hence w ∈ (x, y).
Consequently, there exists h 0 such that w + h, w − h ∈ (x, y). But w = 2−1
(w + h) +
56. From these facts, few as they are, there can be little doubt that
dyeing, and even calico-printing, had made considerable progress
among the ancients; and this could not have taken place without a
considerable knowledge of colouring matters, and of the mordants
by which these colouring matters were fixed. These facts, however,
were probably but imperfectly understood, and could not be the
means of furnishing the ancients with any accurate chemical
knowledge.
VI.—SOAP.
Soap, which constitutes so important and indispensable an
article in the domestic economy of the moderns, was quite unknown
to the ancient inhabitants of Asia, and even of Greece. No allusion to
it occurs in the Old Testament. In Homer, we find Nausicaa, the
daughter of the King of the Phæacians, using nothing but water to
wash her nuptial garments:
They seek the cisterns where Phæacian dames
Wash their fair garments in the limped streams;
Where gathering into depth from falling rills,
The lucid wave a spacious bason fills.
The mules unharness’d range beside the main,
Or crop the verdant herbage of the plain.
Then emulous the royal robes they lave,
And plunge the vestures in the cleansing wave.
Odyssey, vi. 1. 99.
We find, in some of the comic poets, that the Greeks were in
the habit of adding wood-ashes to water to make it a better
detergent. Wood-ashes contain a certain portion of carbonate of
potash, which of course would answer as a detergent; though, from
57. its caustic qualities, it would be injurious to the hands of the
washerwomen. There is no evidence that carbonate of soda, the
nitrum of the ancients, was ever used as a detergent; this is the
more surprising, because we know from Pliny that it was employed
in dyeing, and one cannot see how a solution of it could be
employed by the dyers in their processes without discovering that it
acted powerfully as a detergent.
The word soap (sapo) occurs first in Pliny. He informs us that it
was an invention of the Gauls, who employed it to render their hair
shining; that it was a compound of wood-ashes and tallow, that
there were two kinds of it, hard and soft (spissus et liquidus); and
that the best kind was made of the ashes of the beech and the fat of
goats. Among the Germans it was more employed by the men than
the women.88 It is curious that no allusion whatever is made by
Pliny to the use of soap as a detergent; shall we conclude from this
that the most important of all the uses of soap was unknown to the
ancients?
It was employed by the ancients as a pomatum; and, during the
early part of the government of the emperors, it was imported into
Rome from Germany, as a pomatum for the young Roman beaus.
Beckmann is of opinion that the Latin word sapo is derived from the
old German word sepe, a word still employed by the common people
of Scotland.89
It is well known that the state of soap depends upon the alkali
employed in making it. Soda constitutes a hard soap, and potash a
soft soap. The ancients being ignorant of the difference between the
two alkalies, and using wood-ashes in the preparation of it,
doubtless formed soft soap. The addition of some common salt,
during the boiling of the soap, would convert the soft into hard soap.
As Pliny informs us that the ancients were acquainted both with hard
and soft soap, it is clear that they must have followed some such
process.
58. VII.—STARCH.
The manufacture of starch was known to the ancients. Pliny
informs us that it was made from wheat and from siligo, which was
probably a variety or sub-species of wheat. The invention of starch is
ascribed by Pliny to the inhabitants of the island of Chio, where in
his time the best starch was still made. Pliny’s description of the
method employed by the ancients of making starch is tolerably
exact. Next to the China starch that of Crete was most celebrated;
and next to it was the Egyptian. The qualities of starch were judged
of by the weight; the lightest being always reckoned the best.
VIII.—BEER.
That the ancients were acquainted with wine is universally
known. This knowledge must have been nearly coeval with the origin
of society; for we are informed in Genesis that Noah, after the flood,
planted a vineyard, and made wine, and got intoxicated by drinking
the liquid which he had manufactured.90 Beer also is a very old
manufacture. It was in common use among the Egyptians in the
time of Herodotus, who informs us that they made use of a kind of
wine made from barley, because no vines grew in their country.91
Tacitus informs us, that in his time it was the drink of the
Germans.92 Pliny informs us that it was made by the Gauls, and by
other nations. He gives it the name of cerevisia or cervisia; the name
obviously alluding to the grain from which it was made.
But though the ancients seem acquainted with both wine and
beer, there is no evidence of their having ever subjected these
liquids to distillation, and of having collected the products. This
would have furnished them with ardent spirits or alcohol, of which
59. there is every reason to believe they were entirely ignorant. Indeed,
the method employed by Dioscorides to obtain mercury from
cinnabar, is a sufficient proof that the true process of distillation was
unknown to them. He mixed cinnabar with iron filings, put the
mixture into a pot, to the top of which a cover of stoneware was
luted. Heat was applied to the pot, and when the process was at an
end, the mercury was found adhering to the inside of the cover. Had
they been aware of the method of distilling the quicksilver ore into a
receiver, this imperfect mode of collecting only a small portion of the
quicksilver, separated from the cinnabar, would never have been
practised. Besides, there is not the smallest allusion to ardent spirits,
either in the writings of the poets, historians, naturalists, or medical
men of ancient Greece; a circumstance not to be accounted for had
ardent spirits been known, and applied even to one-tenth of the
uses to which they are put by the moderns.
IX.—STONEWARE.
The manufacture of stoneware vessels was known at a very
early period of society. Frequent allusions to the potter’s wheel occur
in the Old Testament, showing that the manufacture must have been
familiar to the Jewish nation. The porcelain of the Chinese boasts of
a very high antiquity indeed. We cannot doubt that the processes of
the ancients were similar to those of the moderns, though I am not
aware of any tolerably accurate account of them in any ancient
author whatever.
Moulds of plaster of Paris were used by the ancients to take
casts precisely as at present.93
The sand of Puzzoli was used by the Romans, as it is by the
moderns, to form a mortar capable of hardening under water.
60. Pliny gives us some idea of the Roman bricks, which are known
to have been of an excellent quality. There were three sizes of bricks
used by the Romans.
1. Lydian, which were 1½ foot long and 1 foot broad.
2. Tetradoron, which was a square of 16 inches each side.
3. Pentadoron, which was a square, each side of which was 20
inches long.
Doron signifies the palm of the hand: of course it was
equivalent to 4 inches.
X.—PRECIOUS STONES AND MINERALS.
Pliny has given a pretty detailed description of the precious
stones of the ancients; but it is not very easy to determine the
specific minerals to which he alludes.
1. The description of the diamond is tolerably precise. It was
found in Ethiopia, India, Arabia, and Macedonia. But the Macedonian
diamond, as well as the adamas cyprius and siderites, were
obviously not diamonds, but soft stones.
2. The emerald of the ancients (smaragdus) must have varied in
its nature. It was a green, transparent, hard stone; and, as colour
was the criterion by which the ancients distinguished minerals and
divided them into species, it is obvious that very different minerals
must have been confounded together, under the name of emerald.
Sapphire, beryl, doubtless fluor spar when green, and probably even
serpentine, nephrite, and some ores of copper, seem to have
occasionally got the same name. There is no reason to believe that
the emerald of the moderns was known before the discovery of
61. America. At least it has been only found in modern times in America.
Some of the emeralds described by Pliny as losing their colour by
exposure to the sun, must have been fluor spars. There is a
remarkably deep and beautiful green fluor spar, met with some years
ago in the county of Durham, in one of the Weredale mines that
possesses this property. The emeralds of the ancients were of such a
size (13½ feet, large enough to be cut into a pillar), that we can
consider them in no other light than as a species of rock.
3. Topaz of the ancients had a green colour, which is never the
case with the modern topaz. It was found in the island Topazios, in
the Red Sea.94 It is generally supposed to have been the chrysolite
of the moderns. But Pliny mentions a statue of it six feet long. Now
chrysolite never occurs in such large masses. Bruce mentions a
green substance in an emerald island in the Red Sea, not harder
than glass. Might not this be the emerald of the ancients?
4. Calais, from the locality and colour was probably the Persian
turquoise, as it is generally supposed to be.
5. Whether the prasius and chrysoprasius of Pliny were the
modern stones to which these names are given, we have no means
of determining. It is generally supposed that they are, and we have
no evidence to the contrary.
6. The chrysolite of Pliny is supposed to be our topaz: but we
have no other evidence of this than the opinion of M. Du Tems.
7. Asteria of Pliny is supposed by Saussure to be our sapphire.
The lustre described by Pliny agrees with this opinion. The stone is
said to have been very hard and colourless.
8. Opalus seems to have been our opal. It is called, Pliny says,
pæderos by many, on account of its beauty. The Indians called it
sangenon.
62. 9. Obsidian was the same as the mineral to which we give that
name. It was so called because a Roman named Obsidianus first
brought it from Egypt. I have a piece of obsidian, which the late Mr.
Salt brought from the locality specified by Pliny, and which possesses
all the characters of that mineral in its purest state.
10. Sarda was the name of carnelian, so called because it was
first found near Sardis. The sardonyx was also another name for
carnelian.
11. Onyx was a name sometimes given to a rock, gypsum;
sometimes it was a light-coloured chalcedony. The Latin name for
chalcedony was carchedonius, so called because Carthage was the
place where this mineral was exposed to sale. The Greek name for
Carthage was Καρχηδων (carchedon).
12. Carbunculus was the garnet; and anthrax was a name for
another variety of the same mineral.
13. The oriental amethyst of Pliny was probably a sapphire. The
fourth species of amethyst described by Pliny, seems to have been
our amethyst. Pliny derives the name from α (a) and μυθη (mythe),
wine, because it has not quite the colour of wine. But the common
derivation is from α and μυθυω, to intoxicate, because it was used
as an amulet to prevent intoxication.
14. The sapphire is described by Pliny as always opaque, and as
unfit for engraving on. We do not know what it was.
15. The hyacinth of Pliny is equally unknown. From its name it
was obviously of a blue colour. Our hyacinth has a reddish-brown
colour, and a great deal of hardness and lustre.
16. The cyanus of Pliny may have been our cyanite.
17. Astrios agrees very well, as far as the description of Pliny
goes, with the variety of felspar called adularia.
63. 18. Belioculus seems to have been our catseye.
19. Lychnites was a violet-coloured stone, which became electric
by heat. Unless it was a blue tourmalin, I do not know what it could
be.
20. The jasper of the ancients was probably the same as ours.
21. Molochites may have been our malachite. The name comes
from the Greek word μολοχη, mallow, or marshmallow.
22. Pliny considers amber as the juice of a tree concreted into a
solid form. The largest piece of it that he had ever seen weighed 13
lbs. Roman weight, which is nearly equivalent to 9¾ lbs.
avoirdupois. Indian amber, of which he speaks, was probably copal,
or some transparent resin. It may be dyed, he says, by means of
anchusa and the fat of kids.
23. Lapis specularis was foliated sulphate of lime, or selenite.
24. Pyrites had the same meaning among the ancients that it
has among the moderns; at least as far as iron pyrites or bisulphuret
of iron is concerned. Pliny describes two kind of pyrites; namely, the
white (arsenical pyrites), and the yellow (iron pyrites). It was used
for striking fire with steel, in order to kindle tinder. Hence the name
pyrites or firestone.
25. Gagates, from the account given of it by Pliny, was obviously
pit-coal or jet.
26. Marble had the same meaning among the ancients that it
has among the moderns. It was sawed by the ancients into slabs,
and the action of the saw was facilitated by a sand brought for the
purpose from Ethiopia and the isle of Naxos. It is obvious that this
sand was powdered corundum, or emery.
64. 27. Creta was a name applied by the ancients not only to chalk,
but to white clay.
28. Melinum was an oxide of iron. Pliny gives a list of one
hundred and fifty-one species of stones in the order of the alphabet.
Very few of the minerals contained in this list can be made out. He
gives also a list of fifty-two species of stones, whose names are
derived from a fancied resemblance which the stones are supposed
to bear to certain parts of animals. Of these, also, very few can be
made out.
XI.—MISCELLANEOUS OBSERVATIONS.
The ancients seem to have been ignorant of the nature and
properties of air, and of all gaseous bodies. Pliny’s account of air
consists of a single sentence: “Aër densatur nubibus; furit procellis.”
“Air is condensed in clouds, it rages in storms.” Nor is his description
of water much more complete, since it consists only of the following
phrases: “Aquæ subeunt in imbres, rigescunt in grandines,
tumescunt in fluctus, præcipitantur in torrentes.”95 “Water falls in
showers, congeals in hail, swells in waves, and rushes down in
torrents.” In the thirty-eighth chapter of the second book, indeed, he
professes to treat of air; but the chapter contains merely an
enumeration of meteorological phenomena, without once touching
upon the nature and properties of air.
Pliny, with all the philosophers of antiquity, admitted the
existence of the four elements, fire, air, water, and earth; but though
he enumerates these in the fifth chapter of his first book, he never
attempts to explain their nature or properties. Earth, among the
ancients, had two meanings, namely, the planet on which we live,
and the soil upon which vegetables grow. These two meanings still
exist in common language. The meaning afterwards given to the
65. term, earth, by the chemists, did not exist in the days of Pliny, or, at
least, was unknown to him; a sufficient proof that chemistry, in his
time, had made no progress as a science; for some notions
respecting the properties and constituents of those supposed four
elements must have constituted the very foundation of scientific
chemistry.
The ancients were acquainted with none of the acids which at
present constitute so numerous a tribe, except vinegar, or acetic
acid; and even this acid was not known to them in a state of purity.
They knew none of the saline bases, except lime, soda, and potash,
and these very imperfectly. Of course the whole tribe of salts was
unknown to them, except a very few, which they found ready
formed in the earth, or which they succeeded in forming by the
action of vinegar on lead and copper. Hence all that extensive and
most important branch of chemistry, consisting of the combinations
of the acids and bases, on which scientific chemistry mainly
depends, must have been unknown to them.
Sulphur occurring native in large quantities, and being
remarkable for its easy combustibility, and its disagreeable smell
when burning, was known in the very earliest ages. Pliny describes
four kinds of sulphur, differing from each other, probably, merely in
their purity. These were
1. Sulphur vivum, or apyron. It was dug out of the earth
solid, and was doubtless pure, or nearly so. It alone
was used in medicine.
2. Gleba—used only by fullers.
3. Egula—used also by fullers.
Pliny says, it renders woollen stuffs white and soft. It is
obvious from this, that the ancients knew the method
of bleaching flannel by the fumes of sulphur, as
practised by the moderns.
4. The fourth kind was used only for sulphuring matches.
66. Sulphur, in Pliny’s time, was found native in the Æolian islands,
and in Campania. It is curious that he never mentions Sicily, whence
the great supply is drawn for modern manufacture.
In medicine, it seems to have been only used externally by the
ancients. It was considered as excellent for removing eruptions. It
was used also for fumigating.
The word alumen, which we translate alum, occurs often in
Pliny; and is the same substance which the Greeks distinguished by
the name of στυπτηρια (stypteria). It is described pretty minutely by
Dioscorides, and also by Pliny. It was obviously a natural production,
dug out of the earth, and consequently quite different from our
alum, with which the ancients were unacquainted. Dioscorides says
that it was found abundantly in Egypt; that it was of various kinds,
but that the slaty variety was the best. He mentions also many other
localities. He says that, for medical purposes, the most valued of all
the varieties of alumen were the slaty, the round, and the liquid. The
slaty alumen is very white, has an exceedingly astringent taste, a
strong smell, is free from stony concretions, and gradually cracks
and emits long capillary crystals from these rifts; on which account it
is sometimes called trichites. This description obviously applies to a
kind of slate-clay, which probably contained pyrites mixed with it of
the decomposing kind. The capillary crystals were probably similar to
those crystals at present called hair-salt by mineralogists, which
exude pretty abundantly from the shale of the coal-beds, when it
has been long exposed to the air. Hair-salt differs very much in its
nature. Klaproth ascertained by analysis, that the hair-salt from the
quicksilver-mines in Idria is sulphate of magnesia, mixed with a small
quantity of sulphate of iron.96 The hair-salt from the abandoned
coal-pits in the neighbourhood of Glasgow is a double salt,
composed of sulphate of alumina, and sulphate of iron, in definite
proportions; the composition being
1 atom protosulphate of iron,
1½atom sulphate of alumina,
67. 15 atoms water.
I suspect strongly that the capillary crystals from the schistose
alumen of Dioscorides were nearly of the same nature.
From Pliny’s account of the uses to which alumen was applied, it
is quite obvious that it must have varied very much in its nature.
Alumen nigrum was used to strike a black colour, and must therefore
have contained iron. It was doubtless an impure native sulphate of
iron, similar to many native productions of the same nature still met
with in various parts of the world, but not employed; their use
having been superseded by various artificial salts, more definite in
their nature, and consequently more certain in their application, and
at the same time cheaper and more abundant than the native.
The alumen employed as a mordant by the dyers, must have
been a sulphate of alumina more or less pure; at least it must have
been free from all sulphate of iron, which would have affected the
colour of the cloth, and prevented the dyer from accomplishing his
object.97
What the alumen rotundum was, is not easily conjectured.
Dioscorides says, that it was sometimes made artificially; but that
the artificial alumen rotundum was not much valued. The best, he
says, was full of air-bubbles, nearly white, and of a very astringent
taste. It had a slaty appearance, and was found in Egypt or the
Island of Melos.
The liquid alumen was limpid, milky, of an equal colour, free
from hard concretions, and having a fiery shade of colour.98 In its
nature, it was similar to the alumen candidum; it must therefore
have consisted chiefly, at least, of sulphate of alumina.
Bitumen and naphtha were known to the ancients, and used by
them to give light instead of oil; they were employed also as
external applications in cases of disease, and were considered as
68. having the same virtues as sulphur. It is said, that the word
translated salt in the New Testament—“Ye are the salt of the earth:
but if the salt have lost his savour, wherewith shall it be salted? It is
henceforth good for nothing, but to be cast out, and to be trodden
under foot of men”99—it is said, that the word salt in this passage
refers to asphalt, or bitumen, which was used by the Jews in their
sacrifices, and called salt by them. But I have not been able to find
satisfactory evidence of the truth of this opinion. It is obvious from
the context, that the word translated salt could not have had that
meaning among the Jews; because salt never can be supposed to
lose its savour. Bitumen, while liquid, has a strong taste and smell,
which it loses gradually by exposure to the air, as it approaches
more and more to a solid form.
Asphalt was one of the great constituents of the Greek fire. A
great bed of it still existing in Albania, supplied the Greeks with this
substance. Concerning the nature of the Greek fire, it is clear that
many exaggerated and even fabulous statements have been
published. The obvious intention of the Greeks being, probably, to
make their invention as much dreaded as possible by their enemies.
Nitre was undoubtedly one of the most important of its constituents;
though no allusion whatever is ever made. We do not know when
nitrate of potash, the nitre of the moderns, became known in
Europe. It was discovered in the east; and was undoubtedly known
in China and India before the commencement of the Christian era.
The property of nitre, as a supporter of combustion, could not have
remained long unknown after the discovery of the salt. The first
person who threw a piece of it upon a red-hot coal would observe it.
Accordingly we find that its use in fireworks was known very early in
China and India; though its prodigious expansive power, by which it
propels bullets with so great and destructive velocity, is a European
invention, posterior to the time of Roger Bacon.
The word nitre ()רתנ had been applied by the ancients to
carbonate of soda, a production of Egypt, where it is still formed
from sea-water, by some unknown process of nature in the marshes
69. near Alexandria. This is evident, not merely from the account given
of it by Dioscorides and Pliny; for the following passage, from the
Old Testament, shows that it had the same meaning among the
Jews: “As he that taketh away a garment in cold weather, is as
vinegar upon nitre: so is he that singeth songs to a heavy heart.”100
Vinegar poured upon saltpetre produces no sensible effect whatever,
but when poured upon carbonate of soda, it occasions an
effervescence. When saltpetre came to be imported to Europe, it
was natural to give it the same name as that applied to carbonate of
soda, to which both in taste and appearance it bore some faint
resemblance. Saltpetre possessing much more striking properties
than carbonate of soda much more attention was drawn to it, and it
gradually fixed upon itself the term nitre, at first applied to a
different salt. When this change of nomenclature took place does
not appear; but it was completed before the time of Roger Bacon,
who always applies the term nitrum to our nitrate of potash and
never to carbonate of soda.
In the preceding history of the chemical facts known to the
ancients, I have taken no notice of a well-known story related of
Cleopatra. This magnificent and profligate queen boasted to Antony
that she would herself consume a million of sistertii at a supper.
Antony smiled at the proposal, and doubted the possibility of her
performing it. Next evening a magnificent entertainment was
provided, at which Antony, as usual, was present, and expressed his
opinion that the cost of the feast, magnificent as it was, fell far short
of the sum specified by the queen. She requested him to defer
computing till the dessert was finished. A vessel filled with vinegar
was placed before her, in which she threw two pearls, the finest in
the world, and which were valued at ten millions of sistertii; these
pearls were dissolved by the vinegar,101 and the liquid was
immediately drunk by the queen. Thus she made good her boast,
and destroyed the two finest pearls in the world.102 This story,
supposing it true, shows that Cleopatra was aware that vinegar has
the property of dissolving pearls. But not that she knew the nature
70. of these beautiful productions of nature. We now know that pearls
consist essentially of carbonate of lime, and that the beauty is owing
to the thin concentric laminæ, of which they are composed.
Nor have I taken any notice of lime with which the ancients
were well acquainted, and which they applied to most of the uses to
which the moderns put it. Thus it constituted the base of the Roman
mortar, which is known to have been excellent. They employed it
also as a manure for the fields, as the moderns do. It was known to
have a corrosive nature when taken internally; but was much
employed by the ancients externally, and in various ways as an
application to ulcers. Whether they knew its solubility in water does
not appear; though, from the circumstance of its being used for
making mortar, this fact could hardly escape them. These facts,
though of great importance, could scarcely be applied to the rearing
of a chemical structure, as the ancients could have no notion of the
action of acids upon lime, or of the numerous salts which it is
capable of forming. Phenomena which must have remained
unknown till the discovery of the acids enabled experimenters to try
their effects upon limestone and quicklime. Not even a conjecture
appears in any ancient writer that I have looked into, about the
difference between quicklime and limestone. This difference is so
great that it must have been remarked by them, yet nobody seems
ever to have thought of attempting to account for it. Even the
method of burning or calcining lime is not described by Pliny; though
there can be no doubt that the ancients were acquainted with it.
Nor have I taken any notice of leather or the method of tanning
it. There are so many allusions to leather and its uses by the ancient
poets and historians, that the acquaintance of the ancients with it is
put out of doubt. But so far as I know, there is no description of the
process of tanning in any ancient author whatever.
71. CHAPTER III.
CHEMISTRY OF THE ARABIANS.
Hitherto I have spoken of Alchymy, or of the chemical
manufactures of the ancients. The people to whom scientific
chemistry owes its origin are the Arabians. Not that they prosecuted
scientific chemistry themselves; but they were the first persons who
attempted to form chemical medicines. This they did by mixing
various bodies with each other, and applying heat to the mixture in
various ways. This led to the discovery of some of the mineral acids.
These they applied to the metals, c., and ascertained the effects
produced upon that most important class of bodies. Thus the
Arabians began those researches which led gradually to the
formation of scientific chemistry. We must therefore endeavour to
ascertain the chemical facts for which we are indebted to the
Arabians.
When Mahomet first delivered his dogmas to his countrymen
they were not altogether barbarous. Possessed of a copious and
expressive language, and inhabiting a burning climate, their
imaginations were lively and their passions violent. Poetry and fiction
were cultivated by them with ardour, and with considerable success.
But science and inductive philosophy, had made little or no progress
among them. The fatalism introduced by Mahomet, and the blind
enthusiasm which he inculcated, rendered them furious bigots and
determined enemies to every kind of intellectual improvement. The
rapidity with which they overran Asia, Africa, and even a portion of
Europe, is universally known. At that period the western world, was
72. sunk into extreme barbarism, and the Greeks, with whom the
remains of civilization still lingered, were sadly degenerated from
those sages who graced the classic ages. Bent to the earth under
the most grinding but turbulent despotism that ever disgraced
mankind, and having their understandings sealed up by the most
subtle and absurd, and uncompromising superstition, all the energy
of mind, all the powers of invention, all the industry and talent,
which distinguished their ancestors, had completely forsaken them.
Their writers aimed at nothing new or great, and were satisfied with
repeating the scientific facts determined by their ancestors. The
lamp of science fluttered in its socket, and was on the eve of being
extinguished.
Nothing good or great could be expected from such a state of
society. It was, therefore, wisely determined by Providence that the
Mussulman conquerors, should overrun the earth, sweep out those
miserable governors, and free the wretched inhabitants from the
trammels of despotism and superstition. As a despotism not less
severe, and a superstition still more gloomy and uncompromising,
was substituted in their place, it may seem at first sight, that the
conquests of the Mahometans brought things into a worse state than
they found them. But the listless inactivity, the almost deathlike
torpor which had frozen the minds of mankind, were effectually
roused. The Mussulmans displayed a degree of energy and activity
which have few parallels in the history of the world: and after the
conquests of the Mahometans were completed, and the Califs quietly
seated upon the greatest and most powerful throne that the world
had ever seen; after Almanzor, about the middle of the eighth
century, had founded the city of Bagdad, and settled a permanent
and flourishing peace, the arts and sciences, which usually
accompany such a state of society, began to make their appearance.
That calif founded an academy at Bagdad, which acquired much
celebrity, and gradually raised itself above all the other academies in
his dominions. A medical college was established there with powers
to examine all those persons who intended to devote themselves to
73. the medical profession. So many professors and pupils flocked to this
celebrated college, from all parts of the world, that at one time their
number amounted to no fewer than six thousand. Public hospitals
and laboratories were established to facilitate a knowledge of
diseases, and to make the students acquainted with the method of
preparing medicines. It was this last establishment which originated
with the califs that gave a first beginning to the science of chemistry.
In the thirteenth century the calif Mostanser re-established the
academy and the medical college at Bagdad: for both had fallen into
decay, and had been replaced by an infinite number of Jewish
seminaries. Mostanser gave large salaries to the professors,
collected a magnificent library, and established a new school of
pharmacy. He was himself often present at the public lectures.
The successor of Mostanser was the calif Haroun-Al-Raschid, the
perpetual hero of the Arabian tales. He not only carried his love for
the sciences further than his predecessors, but displayed a liberality
and a tolerance for religious opinions, which was not quite consistent
with Mahometan bigotry and superstition. He drew round him the
Syrian Christians, who translated the Greek classics, rewarded them
liberally, and appointed them instructors of his Mahometan subjects,
especially in medicine and pharmacy. He protected the Christian
school of Dschondisabour, founded by the Nestorian Christians,
before the time of Mahomet, and still continuing in a flourishing
state: always surrounded by literary men, he frequently
condescended to take a part in their discussions, and not
unfrequently, as might have been expected from his rank, came off
victorious.
The most enlightened of all the califs was Almamon, who has
rendered his name immortal by his exertions in favour of the
sciences. It was during his reign that the Arabian schools came to be
thoroughly acquainted with Greek science; he procured the
translation of a great number of important works. This conduct
inflamed the religious zeal of the faithful, who devoted him to
74. destruction, and to the divine wrath, for favouring philosophy, and in
that way diminishing the authority of the Koran. Almamon purchased
the ancient classics, from all quarters, and recommended the care of
doing so in a particular manner to his ambassadors at the court of
the Greek emperors. To Leo, the philosopher, he made the most
advantageous offers, to induce him to come to Bagdad; but that
philosopher would not listen to his invitation. It was under the
auspices of this enlightened prince, that the celebrated attempt was
made to determine the size of the earth by measuring a degree of
the meridian. The result of this attempt it does not belong to this
work to relate.
Almotassem and Motawakkel, who succeeded Almamon,
followed his example, favoured the sciences, and extended their
protection to men of science who were Christians. Motawakkel re-
established the celebrated academy and library of Alexandria. But he
acted with more severity than his predecessors with regard to the
Christians, who may perhaps have abused the tolerance which they
enjoyed.
The other vicars of the prophet, in the different Mahometan
states, followed the fine example set them by Almamon. Already in
the eighth century the sovereigns of Mogreb and the western
provinces of Africa showed themselves the zealous friends of the
sciences. One of them called Abdallah-Ebn-Ibadschab rendered
commerce and industry flourishing at Tunis. He himself cultivated
poetry and drew numerous artists and men of science into his state.
At Fez and in Morocco the sciences flourished, especially during the
reign of the Edrisites, the last of whom, Jahiah, a prince possessed
of genius, sweetness, and goodness, changed his court into an
academy, and paid attention to those only who had distinguished
themselves by their scientific knowledge.
But Spain was the most fortunate of all the Mahometan states,
and had arrived at such a degree of prosperity both in commerce,
manufactures, population, and wealth, as is hardly to be credited.
75. The three Abdalrahmans and Alhakem carried, from the eighth to
the tenth century, the country subject to the Calif of Cordova to the
highest degree of splendour. They protected the sciences, and
governed with so much mildness, that Spain was probably never so
happy under the dominion of any Christian prince. Alhakem
established at Cordova an academy, which for several ages was the
most celebrated in the whole world. All the Christians of Western
Europe repaired to this academy in search of information. It
contained, in the tenth century, a library of 280,000 volumes. The
catalogue of this library filled no less than forty-four volumes.
Seville, Toledo, and Murcia, had likewise their schools of science and
their libraries, which retained their celebrity as long as the dominion
of the Moors lasted. In the twelfth century there were seventy public
libraries in that part of Spain which belonged to the Mahometans.
Cordova had produced one hundred and fifty authors, Almeria fifty-
two, and Murcia sixty-two.
The Mahometan states of the east continued also to favour the
sciences. An emir of Irak, Adad-El-Daula by name, distinguished
himself towards the end of the tenth century by the protection which
he afforded to men of science. To him almost all the philosophers of
the age dedicated their works. Another emir of Irak, Saif-Ed-Daula,
established schools at Kufa and at Bussora, which soon acquired
great celebrity. Abou-Mansor-Baharam, established a public library at
Firuzabad in Curdistan, which at its very commencement contained
7000 volumes. In the thirteenth century there existed a celebrated
school of medicine in Damascus. The calif Malek-Adel endowed it
richly, and was often present at the lectures with a book under his
arm.
Had the progress of the sciences among the Arabians been
proportional to the number of those who cultivated them, we might
hail the Saracens as the saviours of literature during the dark and
benighted ages of Christianity; but we must acknowledge with
regret, that notwithstanding the enlightened views of the califs,
notwithstanding the multiplicity of academies and libraries, and the
76. prodigious number of writers, the sciences received but little
improvement from the Arabians. There are very few Arabian writers
in whose works we find either philosophical ideas, successful
researches, new facts, or great and new and important truths. How,
indeed, could such things be expected from a people naturally
hostile to mental exertion; professing a religion which stigmatizes all
exercise of the judgment as a crime, and weighed down by the
heavy yoke of despotism? It was the religion of the Arabians, and
the despotism of their princes, that opposed the greatest obstacles
to the progress of the sciences, even during the most flourishing
period of their civilization.103 Fortunately chemistry was the branch
of science least obnoxious to the religious prejudices of the
Mahometans. It was in it, therefore, that the greatest improvements
were made: of these improvements it will be requisite now to
endeavour to give the reader some idea. Astrology and alchymy,
they both derived from the Greeks: neither of them were
inconsistent with the taste of the nation—neither of them were
anathematized by the Mahometan creed, though Islamism prohibited
magic and all the arts of divination. Alchymy may have suggested
the chemical processes—but the Arabians applied them to the
preparation of medicines, and thus opened a new and most copious
source of investigation.
The chemical writings of the Arabians which I have had an
opportunity of seeing and perusing in a Latin dress, being ignorant
of the original language in which they were written, are those of
Geber and Avicenna.
Geber, whose real name was Abou-Moussah-Dschafar-Al-Soli,
was a Sabean of Harran, in Mesopotamia, and lived during the
eighth century. Very little is known respecting the history of this
writer, who must be considered as the patriarch of chemistry. Golius,
professor of the oriental languages in the University of Leyden,
made a present of Geber’s work in manuscript to the public library.
He translated it into Latin, and published it in the same city in folio,
and afterwards in quarto, under the title of “Lapis
77. Philosophorum.”104 It was translated into English by Richard Russel
in 1678, under the title of, “The Works of Geber, the most famous
Arabian Prince and Philosopher.”105 The works of Geber, so far as
they appeared in Latin or English, consist of four tracts. The first is
entitled, “Of the Investigation or Search of Perfection.” The second is
entitled, “Of the Sum of Perfection, or of the perfect Magistery.” The
third, “Of the Invention of Verity or Perfection.” And the last, “Of
Furnaces, c.; with a Recapitulation of the Author’s Experiments.”
The object of Geber’s work is to teach the method of making
the philosopher’s stone, which he distinguishes usually by the name
of medicine of the third class. The whole is in general written with so
much plainness, that we can understand the nature of the
substances which he employed, the processes which he followed,
and the greater number of the products which he obtained. It is,
therefore, a book of some importance, because it is the oldest
chemical treatise in existence,106 and because it makes us
acquainted with the processes followed by the Arabians, and the
progress which they had made in chemical investigations. I shall
therefore lay before the reader the most important facts contained in
Geber’s work.
1. He considered all the metals as compounds of mercury and
sulphur: this opinion did not originate with him. It is evident from
what he says, that the same notion had been adopted by his
predecessors—men whom he speaks of under the title of the
ancients.
2. The metals with which he was acquainted were gold, silver,
copper, iron, tin, and lead. These are usually distinguished by him
under the names of Sol, Luna, Venus, Mars, Jupiter, and Saturn.
Whether these names of the planets were applied to the metals by
Geber, or only by his translators, I cannot say; but they were always
employed by the alchymists, who never designated the metals by
any other appellations.
78. 3. Gold and silver he considered as perfect metals; but the other
four were imperfect metals. The difference between them depends,
in his opinion, partly upon the proportions of mercury and sulphur in
each, and partly upon the purity or impurity of the mercury and
sulphur which enters into the composition of each.
Gold, according to him, is created of the most subtile substance
of mercury and of most clear fixture, and of a small substance of
sulphur, clean and of pure redness, fixed, clear, and changed from its
own nature, tinging that; and because there happens a diversity in
the colours of that sulphur, the yellowness of gold must needs have
a like diversity.107 His evidence that gold consisted chiefly of
mercury, is the great ease with which mercury dissolves gold. For
mercury, in his opinion, dissolves nothing that is not of its own
nature. The lustre and splendour of gold is another proof of the
great proportion of mercury which it contains. That it is a fixed
substance, void of all burning sulphur, he thinks evident by every
operation in the fire, for it is neither diminished nor inflamed. His
other reasons are not so intelligible.108
Silver, like gold, is composed of much mercury and a little
sulphur; but in the gold the sulphur is red; whereas the sulphur that
goes to the formation of silver is white. The sulphur in silver is also
clean, fixed, and clear. Silver has a purity short of that of gold, and a
more gross inspissation. The proof of this is, that its parts are not so
condensed, nor is it so fixed as gold; for it may be diminished by
fire, which is not the case with gold.109
Iron is composed of earthy mercury and earthy sulphur, highly
fixed, the latter in by far the greatest quantity. Sulphur, by the work
of fixation, more easily destroys the easiness of liquefaction than
mercury. Hence the reason why iron is not fusible, as is the case
with the other metals.110
79. Sulphur not fixed melts sooner than mercury; but fixed sulphur
opposes fusion. What contains more fixed sulphur, more slowly
admits of fusion than what partakes of burning sulphur, which more
easily and sooner flows.111
Copper is composed of sulphur unclean, gross and fixed as to its
greater part; but as to its lesser part not fixed, red, and livid, in
relation to the whole not overcoming nor overcome and of gross
mercury.112
When copper is exposed to ignition, you may discern a
sulphureous flame to arise from it, which is a sign of sulphur not
fixed; and the loss of the quantity of it by exhalation through the
frequent combustion of it, shows that it has fixed sulphur. This last
being in abundance, occasions the slowness of its fusion and the
hardness of its substance. That copper contains red and unclean
sulphur, united to unclean mercury, is, he thinks, evident, from its
sensible qualities.113
Tin consists of sulphur of small fixation, white with a whiteness
not pure, not overcoming but overcome, mixed with mercury partly
fixed and partly not fixed, white and impure.114 That this is the
constitution of tin he thinks evident; for when calcined, it emits a
sulphureous stench, which is a sign of sulphur not fixed: it yields no
flame, not because the sulphur is fixed, but because it contains a
great portion of mercury. In tin there is a twofold sulphur and also a
twofold mercury. One sulphur is less fixed, because in calcining it
gives out a stench as sulphur. The fixed sulphur continues in the tin
after it is calcined. He thinks that the twofold mercury in tin is
evident, from this, that before calcination it makes a crashing noise
when bent, but after it has been thrice calcined, that crashing noise
can no longer be perceived.115 Geber says, that if lead be washed
with mercury, and after its washing melted in a fire not exceeding
the fire of its fusion, a portion of the mercury will remain combined
with the lead, and will give it the crashing noise and all the qualities
80. of tin. On the other hand, you may convert tin into lead. By manifold
repetition of its calcination, and the administration of fire convenient
for its reduction, it is turned into lead.116
Lead, in Geber’s opinion, differs from tin only in having a more
unclean substance commixed of the two more gross substances,
sulphur and mercury. The sulphur in it is burning and more adhesive
to the substance of its own mercury, and it has more of the
substance of fixed sulphur in its composition than tin has.117
Such are the opinions which Geber entertained respecting the
composition of the metals. I have been induced to state them as
nearly in his own words as possible, and to give the reasons which
he has assigned for them, even when his facts were not quite
correct, because I thought that this was the most likely way of
conveying to the reader an accurate notion of the sentiments of this
father of the alchymists, upon the very foundation of the whole
doctrine of the transmutation of metals. He was of opinion that all
the imperfect metals might be transformed into gold and silver, by
altering the proportions of the mercury and sulphur of which they
are composed, and by changing the nature of the mercury and
sulphur so as to make them the same with the mercury and sulphur
which constitute gold and silver. The substance capable of producing
these important changes he calls sometimes the philosopher’s stone,
but generally the medicine. He gives the method of preparing this
important magistery, as he calls it. But it is not worth while to state
his process, because he leaves out several particulars, in order to
prevent the foolish from reaping any benefit from his writings, while
at the same time those readers who possess the proper degree of
sagacity will be able, by studying the different parts of his writings,
to divine the nature of the steps which he omits, and thus profit by
his researches and explanations. But it will be worth while to notice
the most important of his processes, because this will enable us to
judge of the state of chemistry in his time.
81. 4. In his book on furnaces, he gives a description of a furnace
proper for calcining metals, and from the fourteenth chapter of the
fourth part of the first book of his Sum of Perfection, it is obvious
that the method of calcining or oxidizing iron, copper, tin, and lead,
and also mercury and arsenic were familiarly known to him.
He gives a description of a furnace for distilling, and a pretty
minute account of the glass or stoneware, or metallic aludel and
alembic, by means of which the process was conducted. He was in
the habit of distilling by surrounding his aludel with hot ashes, to
prevent it from being broken. He was acquainted also with the
water-bath. These processes were familiar to him. The description of
the distillation of many bodies occurs in his work; but there is not
the least evidence that he was acquainted with ardent spirits. The
term spirit occurs frequently in his writings, but it was applied to
volatile bodies in general, and in particular to sulphur and white
arsenic, which he considered as substances very similar in their
properties. Mercury also he considered as a spirit.
The method of distilling per descensum, as is practised in the
smelting of zinc, was also known to him. He describes an apparatus
for the purpose, and gives several examples of such distillations in
his writings.
He gives also a description of a furnace for melting metals, and
mentions the vessels in which such processes were conducted. He
was acquainted with crucibles; and even describes the mode of
making cupels, nearly similar to those used at present. The process
of cupellating gold and silver, and purifying them by means of lead,
is given by him pretty minutely and accurately: he calls it cineritium,
or at least that is the term used by his Latin translator.
He was in the habit of dissolving salts in water and acetic acid,
and even the metals in different menstrua. Of these menstrua he
nowhere gives any account; but from our knowledge of the
properties of the different metals, and from some processes which
82. he notices, it is easy to perceive what his solvents must have been;
namely, the mineral acids which were known to him, and to which
there is no allusion whatever in any preceding writer that I have had
an opportunity of consulting. Whether Geber was the discoverer of
these acids cannot be known, as he nowhere claims the discovery:
indeed his object was to slur over these acids, as much as possible,
that their existence, or at least their remarkable properties, might
not be suspected by the uninitiated. It was this affectation of secrecy
and mystery that has deprived the earliest chemists of that credit
and reputation to which they would have been justly entitled, had
their discoveries been made known to the public in a plain and
intelligible manner.
The mode of purifying liquids by filtration, and of separating
precipitates from liquids by the same means, was known to Geber.
He called the process distillation through a filter.
Thus the greater number of chemical processes, such as they
were practised almost to the end of the eighteenth century, were
known to Geber. If we compare his works with those of Dioscorides
and Pliny, we shall perceive the great progress which chemistry or
rather pharmacy had made. It is more than probable that these
improvements were made by the Arabian physicians, or at least by
the physicians who filled the chairs in the medical schools, which
were under the protection of the califs: for as no notice is taken of
these processes by any of the Greek or Roman writers that have
come down to us, and as we find them minutely described by the
earliest chemical writers among the Arabians, we have no other
alternative than to admit that they originated in the east.
I shall now state the different chemical substances or
preparations which were known to Geber, or which he describes the
method of preparing in his works.
1. Common salt. This substance occurring in such abundance in
the earth, and being indispensable as a seasoner of food, was
83. known from the earliest ages. But Geber describes the method
which he adopted to free it from impurities. It was exposed to a red
heat, then dissolved in water, filtered, crystallized by evaporation,
and the crystals being exposed to a red heat, were put into a close
vessel, and kept for use.118 Whether the identity of sal-gem (native
salt) and common salt was known to Geber is nowhere said.
Probably not, as he gives separate directions for purifying each.
2. Geber gives an account of the two fixed alkalies, potash and
soda, and gives processes for obtaining them. Potash was obtained
by burning cream of tartar in a crucible, dissolving the residue in
water, filtering the solution, and evaporating to dryness.119 This
would yield a pure carbonate of potash.
Carbonate of soda he calls sagimen vitri, and salt of soda. He
mentions plants which yield it when burnt, points out the method of
purifying it, and even describes the method of rendering it caustic by
means of quicklime.120
3. Saltpetre, or nitrate of potash, was known to him; and Geber
is the first writer in whom we find an account of this salt. Nothing is
said respecting its origin; but there can be little doubt that it came
from India, where it was collected, and known long before
Europeans were acquainted with it. The knowledge of this salt was
probably one great cause of the superiority of the Arabians over
Europeans in chemical knowledge; for it enabled them to procure
nitric acid, by means of which they dissolved all the metals known in
their time, and thus acquired a knowledge of various important
saline compounds, which were of considerable importance.
There is a process for preparing saltpetre artificially, in several
of the Latin copies of Geber, though it does not appear in our English
translation. The method was to dissolve sagimen vitri, or carbonate
of soda, in aqua fortis, to filter and crystallize by evaporation.121 If
this process be genuine, it is obvious that Geber must have been
84. Welcome to our website – the perfect destination for book lovers and
knowledge seekers. We believe that every book holds a new world,
offering opportunities for learning, discovery, and personal growth.
That’s why we are dedicated to bringing you a diverse collection of
books, ranging from classic literature and specialized publications to
self-development guides and children's books.
More than just a book-buying platform, we strive to be a bridge
connecting you with timeless cultural and intellectual values. With an
elegant, user-friendly interface and a smart search system, you can
quickly find the books that best suit your interests. Additionally,
our special promotions and home delivery services help you save time
and fully enjoy the joy of reading.
Join us on a journey of knowledge exploration, passion nurturing, and
personal growth every day!
ebookbell.com