Norms and Inner Products

Norms and Inner Products

Norms and Inner Products Garrett Thomas July 21, 2018 1 About This document is part of a series of notes about math and machine learning. You are free to distribute it as you wish. The latest version can be found at http://gwthomas.github.io/notes. Please report any errors to [email protected]. 2 Norms Norms generalize the notion of length from Euclidean space. A norm on a vector space V is a function k · k : V ! R that satisfies (i) kvk ≥ 0, with equality if and only if v = 0 (ii) kαvk = jαjkvk (iii) ku + vk ≤ kuk + kvk (the triangle inequality) for all u; v 2 V and all α 2 F. A vector space endowed with a norm is called a normed vector space, or simply a normed space. An important fact about norms is that they induce metrics, giving a notion of convergence in vector spaces. Proposition 1. If a vector space V is equipped with a norm k · k : V ! R, then d(u; v) , ku − vk is a metric on V . Proof. The axioms for metrics follow directly from those for norms. If u; v; w 2 V , then (i) d(u; v) = ku − vk ≥ 0, with equality if and only if u − v = 0, i.e. u = v (ii) d(u; v) = ku − vk = k(−1)(v − u)k = j − 1jkv − uk = d(v; u) (iii) d(u; v) = ku − vk = ku − w + w − vk ≤ ku − wk + kw − vk = d(u; w) + d(w; v) so d is a metric as claimed. 1 Therefore any normed space is also a metric space. If a normed space is complete with respect to the distance metric induced by its norm, it is said to be a Banach space. The metric induced by a norm automatically has the property of translation invariance, meaning that d(u + w; v + w) = d(u; v) for any u; v; w 2 V : d(u + w; v + w) = k(u + w) − (v + w)k = ku + w − v − wk = ku − vk = d(u; v) 3 Inner products An inner product on a vector space V over F is a function h·; ·i : V × V ! F satisfying (i) hv; vi ≥ 0, with equality if and only if v = 0 (ii) Linearity in the first slot: hu + v; wi = hu; wi + hv; wi and hαu; vi = αhu; vi (iii) Conjugate symmetry: hu; vi = hv; ui for all u; v; w 2 V and all α 2 F. A vector space endowed with an inner product is called an inner product space. If F = R, the conjugate symmetry condition reduces to symmetry: hu; vi = hv; ui. Additionally, the combination of linearity and conjugate symmetry imply sesquilinearity: hu; αvi = hαv; ui = αhv; ui = αhv; ui = αhu; vi It is also worth noting that h0; vi = 0 for any v because for any u 2 V hu; vi = hu + 0; vi = hu; vi + h0; vi Proposition 2. The function hx; yi , xy is an inner product on C (considered as a vector space over C). Proof. If x; y; z 2 C and α 2 C, then 1. hz; zi = zz = jzj2 ≥ 0, with equality if and only if z = 0 2. hx + y; zi = (x + y)z = xz + yz = hx; zi + hy; zi, and hαx; yi = (αx)y = α(xy) = αhx; yi 3. hx; yi = xy = yx = hy; xi so the function is an inner product as claimed. 3.1 Orthogonality Orthogonality generalizes the notion of perpendicularity from Euclidean space. Two vectors u and v are said to be orthogonal if hu; vi = 0; we write u ? v for shorthand. A set of vectors v1; : : : ; vn is described as pairwise orthogonal (or just orthogonal for short) if vj ? vk for all j 6= k. Proposition 3. If v1; : : : ; vn are nonzero and pairwise orthogonal, then they are linearly indepen- dent. 2 Proof. Assume v1; : : : ; vn are nonzero and pairwise orthogonal, and suppose α1v1 + : : : αnvn = 0. Then for each j = 1; : : : ; n, 0 = h0; vji = hα1v1 + : : : αnvn; vji = α1hv1; vji + ··· + αnhvn; vji = αjhvj; vji Since vj 6= 0, hvj; vji > 0, so αj = 0. Hence v1; : : : ; vn are linearly independent. 4 Norms induced by inner products Any inner product induces a norm given by p kvk , hv; vi Moreover, these norms have certain special properties related to the inner product. The notation k · k is not yet justified as we have not yet shown that this is in fact a norm. We need first a couple of intermediate results, which are useful in their own right. 4.1 Pythagorean Theorem The well-known Pythagorean theorem generalizes naturally to arbitrary inner product spaces. Proposition 4. (Pythagorean theorem) If u ? v, then ku + vk2 = kuk2 + kvk2 Proof. Suppose u ? v, i.e. hu; vi = 0. Then ku + vk2 = hu + v; u + vi = hu; ui + hv; ui + hu; vi + hv; vi = kuk2 + kvk2 as claimed. It is clear that by repeatedly applying the Pythagorean theorem for two vectors, we can obtain a slight generalization: if v1; : : : ; vn are orthogonal, then 2 2 2 kv1 + ··· + vnk = kv1k + ··· + kvnk 4.2 Cauchy-Schwarz inequality The Cauchy-Schwarz inequality is one of the most widespread and useful inequalities in mathe- matics. Proposition 5. If V is an inner product space, then jhu; vij ≤ kukkvk for all u; v 2 V . Equality holds exactly when u and v are linearly dependent. Proof. If v = 0, then equality holds, for jhu; 0ij = 0 = kuk0 = kukk0k. So suppose v 6= 0. In this case we can define hu; vi w v , kvk2 3 which satisfies hu; vi hu; vi hw; u − wi = v; u − v kvk2 kvk2 ! hu; vi hu; vi = hv; ui − hv; vi kvk2 kvk2 hu; vi hv; ui = hv; ui − kvk2 kvk2 kvk2 = 0 Therefore, by the Pythagorean theorem, kuk2 = kw + u − wk2 = kwk2 + ku − wk2 2 hu; vi ≥ v kvk2 2 hu; vi 2 = kvk kvk2 jhu; vij2 = kvk2 which, after multiplying through by kvk2, yields jhu; vij2 ≤ kuk2kvk2 The claimed inequality follows by taking square roots, since both sides are nonnegative. Observe that the only inequality in the reasoning above comes from ku − wk2 ≥ 0, so equality holds if and only if hu; vi 0 = u − w = u − v kvk2 which implies that u and v are linearly dependent. Conversely if u and v are linearly dependent, then u = αv for some α 2 F, and hαv; vi αhv; vi w = v = v = αv kvk2 kvk2 giving u − w = 0, so that equality holds. We now have all the tools needed to prove that k · k is in fact a norm. Proposition 6. If V is an inner product space, then p kvk , hv; vi is a norm on V . Proof. The axioms for norms mostly follow directly from those for inner products, but the triangle inequality requires a bit of work. If u; v 2 V and α 2 F, then (i) kvk = phv; vi ≥ 0 since hv; vi ≥ 0, with equality if and only if v = 0. 4 (ii) kαvk = phαv; αvi = pααhv; vi = pjαj2hv; vi = jαjphv; vi = jαjkvk. (iii) Using the Cauchy-Schwarz inequality, ku + vk2 = hu + v; u + vi = hu; ui + hu; vi + hv; ui + hv; vi = kuk2 + hu; vi + hu; vi + kvk2 = kuk2 + 2 Rehu; vi + kvk2 ≤ kuk2 + 2jhu; vij + kvk2 ≤ kuk2 + 2kukkvk + kvk2 = (kuk + kvk)2 Taking square roots yields ku + vk ≤ kuk + kvk, since both sides are nonnegative. Hence k · k is a norm as claimed. Thus every inner product space is a normed space, and hence also a metric space. If an inner product space is complete with respect to the distance metric induced by its inner product, it is said to be a Hilbert space. 4.3 Orthonormality A set of vectors e1; : : : ; en are said to be orthonormal if they are orthogonal and have unit norm (i.e. kejk = 1 for each j). This pair of conditions can be concisely expressed as hej; eki = δjk, where δjk is the Kronecker delta. Observe that if v 6= 0, then v=kvk is a unit vector in the same \direction" as v (i.e. a positive scalar multiple of v), for v 1 = kvk = 1 kvk kvk Recall that an orthogonal set of nonzero vectors are linearly independent. Unit vectors are necessarily nonzero (since k0k = 0 6= 1), so an orthonormal set of vectors always forms a basis for its span. Not surprisingly, such a basis is referred to as an orthonormal basis. A nice property of orthonormal bases is that vectors' coefficients in terms of this basis can be computed via the inner product. Proposition 7. If e1; : : : ; en is an orthonormal basis for V , then any v 2 V can be written v = hv; e1ie1 + ··· + hv; enien Proof. Since e1; : : : ; en is a basis for V , any v 2 V has a unique expansion v = α1e1 + ··· + αnen where α1; : : : ; αn 2 F. Then for j = 1; : : : ; n, hv; eji = hα1e1 + ··· + αnen; eji = α1he1; eji + ··· + αnhen; eji = αj since hek; eji = 0 for k 6= j and hej; eji = 1. 5 The norm of vectors expressed in an orthonormal basis is also easily found, for 2 2 2 2 2 kα1e1 + ··· + αnenk = kα1e1k + ··· + kαnenk = jα1j + ··· + jαnj by application of the Pythagorean theorem. Combining this with the previous result, we see 2 2 2 kvk = jhv; e1ij + ··· + jhv; enij 4.4 The Gram-Schmidt process The Gram-Schmidt process produces an orthonormal basis for a vector space from an arbitrary basis for the space. Proposition 8. Let v1; : : : ; vn 2 V be linearly independent. Define wj , vj − hvj; e1ie1 − · · · − hvj; ej−1iej−1 wj ej , kwjk for j = 1; : : : ; n.

View Full Text

Details

  • File Type
    pdf
  • Upload Time
    -
  • Content Languages
    English
  • Upload User
    Anonymous/Not logged-in
  • File Pages
    11 Page
  • File Size
    -

Download

Channel Download Status
Express Download Enable

Copyright

We respect the copyrights and intellectual property rights of all users. All uploaded documents are either original works of the uploader or authorized works of the rightful owners.

  • Not to be reproduced or distributed without explicit permission.
  • Not used for commercial purposes outside of approved use cases.
  • Not used to infringe on the rights of the original creators.
  • If you believe any content infringes your copyright, please contact us immediately.

Support

For help with questions, suggestions, or problems, please contact us