Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Advanced Calculus of Several Variables
Advanced Calculus of Several Variables
Advanced Calculus of Several Variables
Ebook794 pages8 hours

Advanced Calculus of Several Variables

Rating: 4 out of 5 stars

4/5

()

Read preview

About this ebook

Modern conceptual treatment of multivariable calculus, emphasizing interplay of geometry and analysis via linear algebra and the approximation of nonlinear mappings by linear ones. Over 400 well-chosen problems. 1973 edition.
LanguageEnglish
Release dateOct 10, 2012
ISBN9780486131955
Advanced Calculus of Several Variables

Read more from C. H. Edwards

Related to Advanced Calculus of Several Variables

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Advanced Calculus of Several Variables

Rating: 4.166666666666667 out of 5 stars
4/5

6 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Advanced Calculus of Several Variables - C. H. Edwards

    Variables

    I

    Euclidean Space and Linear Mappings

    to itself. Multivariable calculus deals in general, and in a somewhat similar way, with mappings from one Euclidean space to another. However a number of new and interesting phenomena appear, resulting from the rich geometric structure of nn.

    n in some detail, as preparation for the development in subsequent chapters of the calculus of functions of an arbitrary number of variables. This generality will provide more clear-cut formulations of theoretical results, and is also of practical importance for applications. For example, an economist may wish to study a problem in which the variables are the prices, production costs, and demands for a large number of different commodities; a physicist may study a problem in which the variables are the coordinates of a large number of different particles. Thus a real-life problem may lead to a high-dimensional mathematical model. Fortunately, modern techniques of automatic computation render feasible the numerical solution of many high-dimensional problems, whose manual solution would require an inordinate amount of tedious computation.

    1THE VECTOR SPACE n

    n is simply the collection of all ordered n-tuples of real numbers. That is,

    Recalling that the Cartesian product A × B of the sets A and B is by definition the set of all pairs (a, bn (n n.

    ³, obtained by identifying the triple (x1, x2, x3) of numbers with that point in space whose coordinates with respect to three fixed, mutually perpendicular coordinate axes are x1, x2, x3 respectively, is familiar to the reader (although we frequently write (x, y, z) instead of (x1, x2, xn in terms of n mutually perpendicular coordinate axes in higher dimensions (however there is a valid question as to what perpendicular means in this general context; we will deal with this in Section 3).

    n are frequently referred to as vectors. Thus a vector is simply an n-tuple of real numbers, and not a directed line segment, or equivalence class of them (as sometimes defined in introductory texts).

    n is endowed with two algebraic operations, called vector addition and scalar multiplication (numbers are sometimes called scalars for emphasis). Given two vectors x = (x1, . . . , xn) and y = (y1, . . . , ynn, their sum x + y is defined by

    , the scalar multiple ax is defined by

    For example, if x = (1, 0, −2, 3) and y = (−2, 1, 4, −5) then x + y = (−1, 1, 2, −2) and 2x = (2, 0, −4, 6). Finally we write 0 = (0, . . . . , 0) and −x = (−1)x, and use x − y as an abbreviation for x + (−y).

    The familiar associative, commutative, and distributive laws for the real numbers imply the following basic properties of vector addition and scalar multiplication:

    (Here x, y, z n, and a and b . For example, to prove V6, let x = (x1, . . . , xn). Then

    The remaining verifications are left as exercises for the student.

    A vector space is a set V together with two mappings V × V → V × V → Vsuch that x + 0 = x such that x + (−x) = 0n is a vector space. For the most part, all vector spaces that we consider will be either Euclidean spaces, or subspaces of Euclidean spaces.

    By a subspace of the vector space V is meant a subset W of V that is itself a vector space (with the same operations). It is clear that the subset W of V is a subspace if and only if it is closed under the operations of vector addition and scalar multiplication (that is, the sum of any two vectors in W is again in W, as is any scalar multiple of an element of W)—properties V1–V8 are then inherited by W from V. Equivalently, W is a subspace of V if and only if any linear combination of two vectors in W is also in W (why?). Recall that a linear combination of the vectors v1, . . . , vk is a vector of the form a1 v1 + · · · + ak vk. The span is the set S of all linear combinations of them, and it is said that S is generated by the vectors v1, . . . , vk.

    Example 1 n is a subspace of itself, and is generated by the standard basis vectors

    since (x1, x2, . . . , xn) = x1 e1 + x2 e2 + · · · + xn enn consisting of the zero vector alone is a subspace, called the trivial n.

    Example 2The n n n−1.

    Example 3Given such that a1 x1 + · · · + an xn n (see Exercise 1.1).

    Example 4The span S n of S, and real numbers r and s

    ³ that are generated by a pair of non-collinear vectors. We will see in the next section that every subspace V n is generated by some finite number, at most n, of vectors; the dimension of the subspace V will be defined to be the minimal number of vectors required to generate Vn of all dimensions between 0 and n ³.

    Example 5If V and W n(the set of all vectors that lie in both V and W). See Exercise 1.2.

    Although most of our attention will be confined to subspaces of Euclidean spaces, it is instructive to consider some vector spaces that are not subspaces of Euclidean spaces.

    Example 6Let . If f + g and af are defined by (f + g)(x) = f(x) + g(x) and (af)(x) = af(xn is the set of all polynomials of degree at most nn which is generated by the polynomials 1, x, x², . . . , xn.

    Exercises

    1.1Verify Example 3.

    1.2n is also a subspace.

    1.3Given subspaces V and W n, denote by V + W the set of all vectors v + w . Show that V + W n.

    1.4If V such that x + 2y = 0 and x + y = 3z, show that V ³.

    1.50 denote the set of all differentiable real-valued functions on [0, 1] such that f(0) = f0 is a vector space, with addition and multiplication defined as in Example 6. Would this be true if the condition f(0) = f(1) = 0 were replaced by f(0) = 0, f(1) = 1?

    1.6Given a set S(S) the set of all real-valued functions on S, that is, all maps S → R(S) is a vector space with the operations defined in ({1, . . . , nn may be regarded as the n-tuple (φ(1), φ(2), . . . , φ(n)).

    2SUBSPACES OF n

    n has precisely n − 1 types of proper subspaces (that is, subspaces other than 0 n itself)—namely, one of each dimension 1 through n − 1.

    In order to define dimension, we need the concept of linear independence. The vectors v1, v2, . . . , vk are said to be linearly independent provided that no one of them is a linear combination of the others; otherwise they are linearly dependent. The following proposition asserts that the vectors v1, . . . , vk are linearly independent if and only if x1 v1 + x2 v2 + · · · + xk vk = 0 implies that x1 = x2 = · · · = xk = 0. For example, the fact that x1 e1 + x2 e2 + · · · + xn en = (x1, x2, . . . , xn) then implies immediately that the standard basis vectors e1, e2, . . . , en n are linearly independent.

    Proposition 2.1 The vectors v1, v2, . . . , vk are linearly dependent if and only if there exist numbers x1, x2, . . . , xk, not all zero, such that x1 v1 + x2 v2 + · · · + xk vk = 0.

    PROOFIf there exist such numbers, suppose, for example, that x1 ≠ 0. Then

    so v1, v2, . . . , vk are linearly dependent. If, conversely, v1 = a2 v2 + · · · + ak vk, then we have x1 v1 + x2 v2 + · · · + xk vk = 0 with x1 = −1 ≠ 0 and xi = ai for i > 1.

    Example 1To show that the vectors x = (1, 1, 0), y = (1, 1, 1), z = (0, 1, 1) are linearly independent, suppose that ax + by + cz = 0. By taking components of this vector equation we obtain the three scalar equations

    Subtracting the first from the second, we obtain c = 0. The last equation then gives b = 0, and finally the first one gives a = 0.

    Example 2The vectors x = (1, 1, 0), y = (1, 2, 1), z = (0, 1, 1) are linearly dependent, because x − y + z = 0.

    It is easily verified (Exercise 2.7) that any two collinear vectors, and any three coplanar vectors, are linearly dependent. This motivates the following definition of the dimension of a vector space. The vector space V has dimension n, dim V = n, provided that V contains a set of n linearly independent vectors, while any n + 1 vectors in V are linearly dependent; if there is no integer n for which this is true, then V is said to be infinite-dimensional. Thus the dimension of a finite-dimensional vector space is the largest number of linearly independent vectors which it contains; an infinite-dimensional vector space is one that contains n linearly independent vectors for every positive integer n.

    Example 3Consider . The functions 1, x, x², . . . , xn are linearly independent because a polynomial a0 + a1x + · · · + an xn is infinite-dimensional.

    One certainly expects the above definition of dimension to imply that Euclidean nn does indeed have dimension n. We see immediately that its dimension is at least n, since it contains the n linearly independent vectors e1, . . . , enn is precisely n, we must prove that any n n are linearly dependent.

    Suppose that v1, . . . , vk are k > n n, and write

    We want to find real numbers x1, . . . , xk, not all zero, such that

    Thus we need to find a nontrivial solution of the homogeneous linear equations

    By a nontrivial solution (x1, x2, . . . , xk) of the system (1) is meant one for which not all of the xi are zero. But k > n, and (1) is a system of n homogeneous linear equations in the k unknowns x1, . . . , xk. (Homogeneous meaning that the right-hand side constants are all zero.)

    It is a basic fact of linear algebra that any system of homogeneous linear equations, with more unknowns than equations, has a nontrivial solution. The proof of this fact is an application of the elementary algebraic technique of elimination of variables. Before stating and proving the general theorem, we consider a special case.

    Example 4Consider the following three equations in four unknowns:

    We can eliminate x1 from the last two equations of (2) by subtracting the first equation from the second one, and twice the first equation from the third one. This gives two equations in three unknowns:

    Subtraction of the first equation of (3) from the second one gives the single equation

    in two unknowns. We can now choose x4 arbitrarily. For instance, if x4 = 1, then x3 = −2. The first equation of (of the system (2).

    The procedure illustrated in this example can be applied to the general case of n equations in the unknowns x1, . . . , xk, k > n. First we order the n equations so that the first equation contains x1, and then eliminate x1 from the remaining equations by subtracting the appropriate multiple of the first equation from each of them. This gives a system of n − 1 homogeneous linear equations in the k − 1 variables x2, . . . , xk. Similarly we eliminate x2 from the last n − 2 of these n − 1 equations by subtracting multiples of the first one, obtaining n − 2 equations in the k − 2 variables x3, x4, . . . , xk. After n − 2 steps of this sort, we end up with a single homogeneous linear equation in the k − n + 1 unknowns xn, xn+1, . . . , xk. We can then choose arbitrary nontrivial values for the extra variables xn+1, xn+2, . . . , xk (such as xn+1 = 1, xn+2 = · · · = xk = 0), solve the final equation for xn, and finally proceed backward to solve successively for each of the eliminated variables xn−1, xn−2, . . . , x1. The reader may (if he likes) formalize this procedure to give a proof, by induction on the number n of equations, of the following result.

    Theorem 2.2If k > n, then any system of n homogeneous linear equations in k unknowns has a nontrivial solution.

    By the discussion preceding n = n.

    Corollary 2.3 Any n n are linearly dependent.

    We have seen that the Linearly Independent vectors e1, e2, . . . , en n. A set of linearly independent vectors that generates the vector space V is called a basis for V. Since x = (x1, x2, . . . , xn) = x1e1 + x2 e2 + · · · + xn en, it is clear that the basis vectors e1, . . . , en generate V uniquely; that is, if x = y1e1 + y2 e2 + · · · + yn en also, then xi = yi for each in can be expressed in one and only one way as a linear combination of e1, . . . , en. Any set of n linearly independent vectors in an n-dimensional vector space has this property.

    Theorem 2.4If the vectors v1, . . . , vn in the n-dimensional vector space V are linearly independent, then they constitute a basis for V, and furthermore generate V uniquely.

    , the vectors v, v1, . . . , vn are linearly dependent, so by Proposition 2.1 there exist numbers x, x1, . . . , xn, not all zero, such that

    If x = 0, then the fact that v1, . . . , vn are linearly independent implies that x1 = · · · = xn = 0. Therefore x ≠ 0, so we solve for v:

    Thus the vectors v1, . . . , vn generate V, and therefore constitute a basis for V. To show that they generate V uniquely, suppose that

    Then

    So, since v1, . . . , vn are linearly independent, it follows that ai ai′ = 0, or ai = ai′, for each i.

    n has a basis which contains fewer than n elements. But the following theorem shows that this cannot happen.

    Theorem 2.5If dim V = n, then each basis for V consists of exactly n vectors.

    PROOFLet w1, w2, . . . , wn be n linearly independent vectors in V. If there were a basis v1, v2, . . . , vm for V with m < n, then there would exist numbers {aij} such that

    Since m < n, Theorem 2.2 supplies numbers x1, . . . , xn not all zero, such that

    But this implies that

    which contradicts the fact that w1, . . . , wn are linearly independent. Consequently no basis for V can have m < n elements.

    n. If V nby (Corollary 2.3, and if k = n, then V n by Theorem 2.4. If k > 0, then any k linearly independent vectors in V generate V, and no basis for V contains fewer than k vectors Theorem 2.5).

    Exercises

    2.1Why is it true that the vectors v1, . . . , vk are linearly dependent if any one of them is zero? If any subset of them is linearly dependent?

    2.2n ?

    (a)(1, 0) and (1, 1).

    (b)(1, 0, 0), (1, 1, 0), and (0, 0, 1).

    (c)(1, 1, 1), (1, 1, 0), and (1, 0, 0).

    (d)(1, 1, 1, 0), (1, 0, 0, 0), (0, 1, 0, 0), and (0, 0, 1, 0).

    (e)(1, 1, 1, 1), (1, 1, 1, 0), (1, 1, 0, 0), and (1, 0, 0, 0).

    2.3Find the dimension of the subspace V ⁴ that is generated by the vectors (0, 1, 0, 1), (1, 0, 1, 0), and (1, 1, 1, 1).

    2.4Show that the vectors (1, 0, 0, 1), (0, 1, 0, 1), (0, 0, 1, 1) form a basis for the subspace V ⁴ which is defined by the equation x1 + x2 + x3 − x4 = 0.

    2.5Show that any set v1, . . . , vk, of linearly independent vectors in a vector space V can be extended to a basis for V. That is, if k < n = dim V, then there exist vectors vk + 1, . . . , vn in V such that v1, . . . vn is a basis for V.

    2.6Show that Theorem 2.5 is equivalent to the following theorem: Suppose that the equations

    have only the trivial solution x1 = · · · = xn = 0. Then, for each b = (b1, . . . , bn), the equations

    have a unique solution. Hint: Consider the vectors aj = (a1j, a2j, . . . , anj), j = 1, . . . , n.

    2.7Verify that any two collinear vectors, and any three coplanar vectors, are linearly dependent.

    3INNER PRODUCTS AND ORTHOGONALITY

    n n with an inner product. An inner (scalar) product on the vector space V is a function V × V , which associates with each pair (x, y) of vectors in V x, y , and satisfies the following three conditions:

    The third of these conditions is linearity in the first variable; symmetry then gives linearity in the second variable also. Thus an inner product on V is simply a positive, symmetric, bilinear function on V × V0, 0 = 0 (see Exercise 3.1).

    The usual inner product n is denoted by x · y and is defined by

    where x = (x1, . . . , xn), y = (y1, . . . , yn). It should be clear that this definition satisfies conditions (n see Example 2 below), but we shall use only the usual one.

    Example 1Denote the vector space of all continuous functions on the interval [a, b], and define

    It is obvious that this definition satisfies conditions SP2 and SP3. It also satisfies SP1, because if f(t0) ≠ 0, then by continuity (f(t))² > 0 for all t in some neighborhood of t0, so

    .

    Example 2Let a, b, c be real numbers with a > 0, ac − b² > 0, so that the quadratic form q(x) = ax1² + 2bx1x2 + cx2² is positive-definite (see x, y = ax1y1 + bx1y2 + bx2y1 + cx2y² (why?). With a = c = 1, b ².

    An inner product on the vector space V , called its norm x . In general, a norm on the vector space V is a real-valued function x x on V satisfying the following conditions:

    0 = 0.

    on V is defined by

    It is clear that SP1–SP3 and this definition imply conditions N1 and N2, but the triangle inequality is not so obvious; it will be verified below.

    n is the Euclidean norm

    nnx will denote the Euclidean norm unless otherwise specified.

    Example 3 x xxn }, the maximum of the absolute values of the coordinates of x, n (see Exercise 3.2).

    Example 4 x xxxn n (again see Exercise 3.2).

    A norm on V provides a definition of the distance d(x, y) between any two points x and y of V:

    Note that a distance function d defined in this way satisfies the following three conditions:

    for any three points x, y, z. Conditions D1 and D2 follow immediately from N1 and N2, respectively, while

    by N3. Figure 1.1 indicates why N3 (or D3) is referred to as the triangle inequality.

    Figure 1.1

    The distance function that comes in this way from the Euclidean norm is the familiar Euclidean distance function

    Thus far we have seen that an inner product on the vector space V yields a norm on V, which in turn yields a distance function on V, except that we have not yet verified that the norm associated with a given inner product does indeed satisfy the triangle inequality. The triangle inequality will follow from the Cauchy–Schwarz inequality of the following theorem.

    Theorem 3.1is an inner product on a vector space V, then

    [where the norm is the one defined by (2)].

    PROOFThe inequality is trivial if either x or y is zero, so assume neither is. If u = xx and v = yy u v = 1. Hence

    or

    Replacing x by −x, we obtain

    also, so the inequality follows.

    n, it takes the form

    , with the inner product of Example 1, it becomes

    note that

    x, y = 0, in which case x and y are perpendicular (see the definition below), then the second equality in the above proof gives

    This is the famous theorem associated with the name of Pythagoras (Fig. 1.2).

    Figure 1.2

    Recalling the formula x · y x y cos θ ², we are motivated to define the angle (x, yby

    by the Cauchy–Schwarz inequality. In particular we say that x and y are orthogonal (or perpendicular) if and only if x · y (x, y) = arccos π/2 = 0.

    A set of nonzero vectors v1, v2, . . . in V is said to be an orthogonal set if

    whenever i ≠ j. If in addition each vi vi, vi = 1, then the set is said to be orthonormal.

    Example 5The standard basis vectors e1, . . . , en n.

    Example 6The (infinite) set of functions

    (see Example 1 and Exercise 3.11). This fact is the basis for the theory of Fourier series.

    The most important property of orthogonal sets is given by the following result.

    Theorem 3.2Every finite orthogonal set of nonzero vectors is linearly independent.

    PROOFSuppose that

    Taking the inner product with vi, we obtain

    vi, vi = 0 for i ≠ j if the vectors v1, . . . , vk vi, vi ≠ 0, so ai = 0. Thus (3) implies a1 = · · · = ak = 0, so the orthogonal vectors v1, . . . , vk are linearly independent.

    We now describe the important Gram–Schmidt orthogonalization process for constructing orthogonal bases. It is motivated by the following elementary construction. Given two linearly independent vectors v and w1, we want to find a nonzero vector w2 that lies in the subspace spanned by v and w1, and is orthogonal to w1. Figure 1.3 suggests that such a vector w2 can be obtained by subtracting from v an appropriate multiple cw1 of w1. To determine c,

    Figure 1.3

    w1, v cw= 0 for c v, ww1, w. The desired vector is therefore

    obtained by subtracting from v the "component of v parallel to ww2, w= 0, while w2 ≠ 0 because v and w1 are linearly independent.

    Theorem 3.3If V is a finite-dimensional vector space with an inner product, then V has an orthogonal basis.

    n has an orthogonal basis.

    PROOFWe start with an arbitrary basis v1, . . . , vn for V. Let w1 = v1. Then, by the preceding construction, the nonzero vector

    is orthogonal to w1 and lies in the subspace generated by v1 and v2.

    Suppose inductively that we have found an orthogonal basis w1, . . . , wk for the subspace of V that is generated by v1, . . . , vk. The idea is then to obtain wk+1 by subtracting from vk+1 its components parallel to each of the vectors w1, . . . , wk. That is, define

    where ci vk+1, wi wi, wi wk+1, wi vk+1, wi ci wi, wi , and wk+1 ≠ 0, because otherwise vk+1 would be a linear combination of the vectors w1, . . . , wk, and therefore of the vectors v1, . . . , vk. It follows that the vectors w1, . . . , wk+1 form an orthogonal basis for the subspace of V that is generated by v1, . . . , vk+1.

    After a finite number of such steps we obtain the desired orthogonal basis w1, . . . , wn for V.

    It is the method of proof of Theorem 3.3 that is known as the Gram–Schmidt orthogonalization process, summarized by the equations

    defining the orthogonal basis w1, . . . , wn in terms of the original basis v1, . . . , vn.

    Example 7To find an orthogonal basis for the subspace V ⁴ spanned by the vectors v1 = (1, 1, 0, 0), v2 = (1, 0, 1, 0), v3 = (0, 1, 0, 1), we write

    Example 8Let denote the vector space of polynomials in x, with inner product defined by

    By applying the Gram–Schmidt orthogonalization process to the linearly independent elements 1, x, x², . . . , xn, the first five elements of which are

    (see Exercise 3.12). Upon multiplying the polynomials {pn(x)} by appropriate constants, one obtains the famous Legendre polynomials

    etc.

    can be expressed as a linear combination of orthogonal basis vectors w1, . . . , wn for V. Writing

    and taking the inner product with wi, we immediately obtain

    so

    This is especially simple if w1, . . . , wn is an orthonormal basis for V:

    Of course orthonormal basis vectors are easily obtained from orthogonal ones, simply by dividing by their lengths. In this case the coefficient v · wi of wi in (5) is sometimes called the Fourier coefficient of v with respect to wicorresponding to the orthogonal functions of Example 6 are

    Writing

    by

    and

    It can then be established, under appropriate conditions on f, that the infinite series

    converges to f(x). This infinite series may be regarded as an infinite-dimensional analog of (5).

    Given a subspace V n, denote by V n, each of which is orthogonal to every vector in V. Then it is easy to show that V n, called the orthogonal complement of V (Exercise 3.3). The significant fact about this situation is that the dimensions add up as they should.

    Theorem 3.4If V n, then

    PROOF By Theorem 3.3, there exists an orthonormal basis v1, . . . , vr for V, and an orthonormal basis w1, . . . , ws for V . Then the vectors v1, . . . , vr, w1, . . . , ws are orthornormal, and therefore linearly independent. So in order to conclude from Theorem 2.5 that r + s = n n, define

    Then y · vi = x · vi − (x · vi)(vi · vi) = 0 for each i = 1, . . . , r. Since y is orthogonal to each element of a basis for V(Exercise 3.4). Therefore Eq. (5) above gives

    This and (7) then yield

    so the vectors v1, . . . , vr, w1, . . . , ws n.

    Example 9Consider the system

    homogeneous linear equations in x1, . . . , xn. If ai = (ai1, . . . , ain), i = 1, . . . , k, then these equations can be rewritten as

    Therefore the set S of all solutions of (that are orthogonal to the vectors a1, . . . , ak. If V n generated by a1, . . . , ak, it follows that S = V (Exercise 3.4). If the vectors a1, . . . , ak are linearly independent, we can then conclude from Theorem 3.4 that dim S = n − k.

    Exercises

    3.10, 0 = 0.

    3.2Verify that the functions defined in n.

    3.3If V n, prove that V is also a subspace.

    3.4If the vectors a1, . . . , ak generate the subspace V n.

    3.5

    3.6Let a1, a2, . . . , an n. If x = s1a1 + · · · + snan and y = t1a1 + · · · + tnan, show that x · y = s1t1 + · · · + sn tn. That is, in computing x · y, one may replace the coordinates of x and y n.

    3.7⁴.

    3.8Orthogonalize the basis

    n.

    3.9Find an orthogonal basis for the 3-dimensional subspace V ⁴ that consists of all solutions of the equation x1 + x2 + x3

    Enjoying the preview?
    Page 1 of 1