logo资料库

《The Elements of Statistical Learning》教材答案.pdf

第1页 / 共121页
第2页 / 共121页
第3页 / 共121页
第4页 / 共121页
第5页 / 共121页
第6页 / 共121页
第7页 / 共121页
第8页 / 共121页
资料共121页,剩余部分请下载后查看
A Solution Manual and Notes for: The Elements of Statistical Learning by Jerome Friedman, Trevor Hastie, and Robert Tibshirani John L. Weatherwax∗ David Epstein† 5 March 2018 Introduction The Elements of Statistical Learning is an influential and widely studied book in the fields of machine learning, statistical inference, and pattern recognition. It is a standard recom- mended text in many graduate courses on these topics. It is also very challenging, particularly if one faces it without the support of teachers who are expert in the subject matter. For various reasons, both authors of these notes felt the need to understand the book well, and therefore to produce notes on the text when we found the text difficult at first reading, and answers to the exercises. Gaining understanding is time-consuming and intellectually demanding, so it seemed sensible to record our efforts in LaTeX, and make it available on the web to other readers. A valuable by-product of writing up mathematical material, is that often one finds gaps and errors in what one has written. Now it is well-known in all branches of learning, but in particular in mathematical learning, that the way to learn is to do, rather than to read. It is all too common to read through some material, convince oneself that one understands it well, but then find oneself at sea when trying to apply it in even a slightly different situation. Moreover, material understood only at a shallow level is easily forgotten. It is therefore our strong recommendation that readers of the book should not look at our responses to any of the exercises before making a substantial effort to understand it without this aid. Any time spent in this way, even if it ends without success, will make our solutions ∗wax@alum.mit.edu †David.Epstein@warwick.ac.uk 1
easier to understand and more memorable. Quite likely you will find better solutions than those provided here—let us know when you do! As far as teachers of courses based on the text are concerned, our notes may be regarded as a disadvantage, because it may be felt that the exercises in the book can no longer be used as a source of homework or questions for tests and examinations. This is a slight conflict of interest between teachers of courses on the one hand, and independent learners on the other hand. Finding ourselves in the ranks of the independent learners, the position we take is hardly surprising. Teachers of courses might benefit by comparing their own solutions with ours, as might students in classes and independent learners. If you, the reader, find a problem difficult and become stuck, our notes might enable you to unstick yourself. In addition, there is a myriad of materials on any of the topics this book covers. A search for “statistical learning theory” on Google (as of 1 January 2010) gave over 4 million hits. The possible disadvantage of not being able to use the book’s problems in an academic course is not really such a large one. Obtaining additional supplementary problems at the right level for an academic course is simply a matter of a visit to the library, or a little time spent surfing the net. And, of course, although one is not supposed to say this, teachers of courses can also get stuck. Acknowledgments We are hoping that people will find errors, offer insightful suggestions, and generally improve the understanding of the text, the exercises and of our notes, and that they write to us about this. We will incorporate additions that we think are helpful. Our plan is to gradually add material as we work through the book. Comments and criticisms are always welcome. Special thanks to (most recent comments are listed first): Nicola Doninelli, Mark-Jan Nederhof for solutions in Chapter 5. Dan Wang for his bug report in the AdaBoost code, Liuzhou Zhuo for his comments on Exercise 3.25 and Ruchi Dhiman for his comments on Chapter 4. We use the numbering found in the on-line (second edition) version of this text. The first edition should be similar but may have some differences. 2
Chapter 2 (Overview of Supervised Learning) Statistical Decision Theory We assume a linear model: that is we assume y = f (x) + ε, where ε is a random variable with mean 0 and variance σ2, and f (x) = xT β. Our expected predicted error (EPE) under the squared error loss is EPE(β) =Z (y − xT β)2Pr(dx, dy) . (1) We regard this expression as a function of β, a column vector of length p + 1. In order to find the value of β for which it is minimized, we equate to zero the vector derivative with respect to β. We have ∂EPE ∂β =Z 2 (y − xT β) (−1)x Pr(dx, dy) = −2Z (y − xT β)xPr(dx, dy) . (2) Now this expression has two parts. The first has integrand yx and the second has integrand (xT β)x. Before proceeding, we make a quick general remark about matrices. Suppose that A, B and C are matrices of size 1 × p matrix, p × 1 and q × 1 respectively, where p and q are positive integers. Then AB can be regarded as a scalar, and we have (AB)C = C(AB), each of these expressions meaning that each component of C is multiplied by the scalar AB. If q > 1, the expressions BC, A(BC) and ABC are meaningless, and we must avoid writing them. On the other hand, CAB is meaningful as a product of three matrices, and the result is the q × 1 matrix (AB)C = C(AB) = CAB. In our situation we obtain (xT β)x = xxT β. From ∂EPE/∂β = 0 we deduce for the particular value of β that minimizes the EPE. Since this value of β is a constant, it can be taken out of the expectation to give E[yx] − E[xxT β] = 0 (3) which gives Equation 2.16 in the book. β = E[xxT ]−1E[yx] , We now discuss some points around Equations 2.26 and 2.27. We have ˆβ = (X T X)−1X T y = (X T X)−1X T (Xβ + ε) = β + (X T X)−1X T ε. So This immediately gives ˆy0 = xT 0 ˆβ = xT 0 β + xT 0 (X T X)−1X T ε. (4) (5) ˆy0 = xT 0 β + ℓi(x0)εi NXi=1 3
where ℓi(x0) is the i-th element of the N-dimensional column vector X(X T X)−1x0, as stated at the bottom of page 24 of the book. We now consider Equations 2.27 and 2.28. The variation is over all training sets T , and over all values of y0, while keeping x0 fixed. Note that x0 and y0 are chosen independently of T and so the expectations commute: Ey0|x0ET = ET Ey0|x0. Also ET = EX EY|X . We write y0 − ˆy0 as the sum of three terms y0 − xT 0 β − (ˆy0 − ET (ˆy0)) −ET (ˆy0) − xT 0 β = U1 − U2 − U3. In order to prove Equations 2.27 and 2.28, we need to square the expression in Equation 6 and then apply various expectation operators. First we consider the properties of each of the three terms, Ui in Equation 6. We have Ey0|x0U1 = 0 and ET U1 = U1ET . Despite the notation, ˆy0 does not involve y0. So Ey0|x0U2 = U2Ey0|x0 and clearly ET U2 = 0. Equation 5 gives (6) (7) since the expectation of the length N vector ε is zero. This shows that the bias U3 is zero. U3 = ET (ˆy0) − xT 0 β = xT 0 EX(X T X)−1X T EY|X ε = 0 We now square the remaining part of Equation 6 and then then apply Ey0|x0ET . The cross- term U1U2 gives zero, since Ey0|x0(U1U2) = U2Ey0|x0(U1) = 0. (This works in the same way if ET replaces Ey0|x0.) We are left with two squared terms, and the definition of variance enables us to deal imme- diately with the first of these: Ey0|x0ET U 2 1 = Var(y0|x0) = σ2. It remains to deal with the term ET (ˆy0 − ET ˆy0)2 = VarT (ˆy0) in Equation 2.27. Since the bias U3 = 0, we know that ET ˆy0 = xT 0 β. If m is the 1× 1-matrix with entry µ, then mmT is the 1× 1-matrix with enty µ2. It follows from Equation 5 that the variance term in which we are interested is equal to Since ET = EX EY|X , and the expectation of εεT is σ2IN , this is equal to ET xT 0 (X T X)−1X T εεT X(X T X)−1x0 . σ2xT 0 ET (X T X)−1 x0 = σ2xT 0 EXX T X/N−1 x0/N. (8) We suppose, as stated by the authors, that the mean of the distribution giving rise to X and x0 is zero. For large N, X T X/N is then approximately equal to Cov(X) = Cov(x0), the p × p-matrix-variance-covariance matrix relating the p components of a typical sample vector x—as far as EX is concerned, this is a constant. Applying Ex0 to Equation 8 as in Equation 2.28, we obtain (approximately) σ2Ex0xT 0 Cov(X)−1x0 /N = σ2Ex0tracexT 0 Cov(X)−1x0 /N 0 /N = σ2Ex0traceCov(X)−1x0xT = σ2traceCov(X)−1Cov(x0) /N = σ2trace(Ip)/N = σ2p/N. (9) 4
This completes our discussion of Equations 2.27 and 2.28. Notes on Local Methods in High Dimensions The most common error metric used to compare different predictions of the true (but un- known) mapping function value f (x0) is the mean square error (MSE). The unknown in the above discussion is the specific function mapping function f (·) which can be obtained via different methods many of which are discussed in the book. In supervised learning to help with the construction of an appropriate prediction ˆy0 we have access to a set of “training samples” that contains the notion of randomness in that these points are not under complete control of the experimenter. One could ask the question as to how much square error at our predicted input point x0 will have on average when we consider all possible training sets T . We can compute, by inserting the “expected value of the predictor obtained over all training sets”, ET (ˆy0) into the definition of quadratic (MSE) error as MSE(x0) = ET [f (x0) − ˆy0]2 = ET [ˆy0 − ET (ˆy0) + ET (ˆy0) − f (x0)]2 = ET [(ˆy0 − ET (ˆy0))2 + 2(ˆy0 − ET (ˆy0))(ET (ˆy0) − f (x0)) + (ET (ˆy0) − f (x0))2] = ET [(ˆy0 − ET (ˆy0))2] + (ET (ˆy0) − f (x0))2 . Where we have expanded the quadratic, distributed the expectation across all terms, and noted that the middle term vanishes since it is equal to ET [2(ˆy0 − ET (ˆy0))(ET (ˆy0) − f (x0))] = 0 , because ET (ˆy0) − ET (ˆy0) = 0. and we are left with MSE(x0) = ET [(ˆy0 − ET (ˆy0))2] + (ET (ˆy0) − f (x0))2 . (10) The first term in the above expression ET [(ˆy0 − ET (ˆy0))2] is the variance of our estimator ˆy0 and the second term (ET (ˆy0) − f (x0))2 is the bias (squared) of our estimator. This notion of variance and bias with respect to our estimate ˆy0 is to be understood relative to possible training sets, T , and the specific computational method used in computing the estimate ˆy0 given that training set. Exercise Solutions Ex. 2.1 (target coding) The authors have suppressed the context here, making the question a little mysterious. For example, why use the notation ¯y instead of simply y? We imagine that the background is something like the following. We have some input data x. Some algorithm assigns to x the probability yk that x is a member of the k-th class. This would explain why we are told to assume that the sum of the yk is equal to one. (But, if that reason is valid, then we should 5
also have been told that each yk ≥ 0.) In fact, neither of these two assumptions is necessary to provide an answer to the question. The hyphen in K-classes seems to be a misprint, and should be omitted. We restate the question, clarifying it, but distancing it even further from its origins. Let K > 0 be an integer. For each k with 1 ≤ k ≤ K, let tk be the K-dimensional vector that has 1 in the k-th position and 0 elsewhere. Then, for any K-dimensional vector y, the k for which yk is largest coincides with the k for which tk is nearest to y. By expanding the quadratic we find that argmink||y − tk|| = argmink||y − tk||2 = argmink = argmink = argmink (yi − (tk)i)2 KXi=1 KXi=1(yi)2 − 2yi(tk)i + (tk)i KXi=1−2yi(tk)i + (tk)i 2 , 2 since the sum PK PK k=1 (tk)i i=1 yi 2 = 1. AlsoP yi(tk)i = yk. This means that 2 is the same for all classes k. Notice that, for each k, the sum argmink||y − tk|| = argmink (−2yk + 1) = argmink(−2yk) = argmaxkyk . Ex. 2.2 (the oracle revealed) For this problem one is supposed to regard the points pi and qi below as fixed. If one does not do this, and instead averages over possible choices, then since the controlling points are (1,0) and (0,1), and all probabilities are otherwise symmetric between these points when one integrates out all the variation, the answer must be that the boundary is the perpendicular bisector of the interval joining these two points. The simulation draws 10 points p1, . . . , p10 ∈ R2 from N 1 R2 from N 0 0 , I2 and 10 points q1, . . . , q10 ∈ 1 , I2. The formula for the Bayes decision boundary is given by equating likelihoods. We get an equation in the unknown z ∈ R2, giving a curve in the plane: Xi exp(−5kpi − zk2/2) =Xj exp(−5kqj − zk2/2) . 6
In this solution, the boundary is given as the equation of equality between the two proba- bilities, with the pi and qj constant and fixed by previously performed sampling. Each time one re-samples the pi and qj, one obtains a different Bayes decision boundary. Ex. 2.3 (the median distance to the origin) We denote the N-tuple of data points by (x1, . . . , xN ). Let ri = kxik. Let U(A) be the set of all N-tuples with A < r1 < . . . < rN < 1. Ignoring subsets of measure zero, the set of all N-tuples is the disjoint union of the N! different subsets obtained from U(0) by permuting the indexing set (1, . . . , N). We will look for A > 0 such that the measure of U(A) is half the measure of U(0). The same A will work for each of our N! disjoint subsets, and will therefore give the median for the distance of the smallest xi from the origin. We want to find A such that ZU (A) dx1 . . . dxN = 1 2ZU (0) dx1 . . . dxN . We convert to spherical coordinates. Since the coordinate in the unit sphere Sp−1 contributes the same constant on each side of the equality, obtaining ZA 0, which is an N-dimensional simplex scaled down by a factor (1−Ap). The right-hand integral is dealt with in the same way, setting A = 0. Since the region of integration is N-dimensional, the measure is multiplied by (1 − Ap)N . We solve for A by solving (1 − Ap)N = 1/2. We obtain A = (1 − 2−1/N )1/p, as required. 7
Ex. 2.4 (projections aT x are distributed as normal N(0, 1)) The main point is thatPkxik2 is invariant under the orthogonal group. As a consequence the standard normal distribution exists on any finite dimensional inner product space (by fixing an orthonormal basis). Further, if Rp is written as the orthogonal sum of two vector subspaces, then the product of standard normal distributions on each of the subspaces gives the standard normal distribution on Rp. Everything else follows from this. The on-line edition is correct except that √10 ≈ 3.162278. So, the figure given should be 3.2 instead of 3.1. Note that the version of this question posed in the first edition makes incorrect claims. The first edition talks of the “center of the training points” and the on-line edition talks of the “origin”. The answers are very different. This is shown up best by taking only one training point. Ex. 2.5 (the expected prediction error under least squares) Part (a): In order to discuss Equation (2.27), we go back to (2.26) on page 24. We have y = Xβ + ε, where y and ε are N × 1, X is N × p, and β is p × 1. Hence ˆβ = (X T X)−1X T y = β + (X T X)−1X T ε. Since X and ε are independent variables, ET (ε) = 0 and ET (εT ε) = σ2IN , we have ET ( ˆβ) = β and VarT ( ˆβ) = ET ( ˆβT β) − ET ( ˆβ)ET ( ˆβT ) = (X T X)−1σ2. Now we prove (2.27) on page 26. Note that y0 is constant for the distribution T . Note also ˆβ does not depend on y0 and so the same is true for that, if x0 is held constant, ˆy0 = xT 0 ET (ˆy0) and VarT (ˆy0). Let u = Ey0|x0(y0) = xT 0 β. Then ET (y0 − ˆy0)2 =y2 0 − u2 +ET (ˆy2 0) − (ET ˆy0)2 +(ET ˆy0)2 − 2y0ET ˆy0 + u2 . Therefore We have Ey0|x0ET (y0 − ˆy0)2 = Var(y0|x0) + VarT (ˆy0) + (ET (ˆy0) − u)2 . ET (ˆy0) = xT 0 ET ( ˆβ) = xT 0 β = u. Part (b): If A is a p×q matrix and B is a q×p matrix, then it is standard that trace(AB) = 0 (X T X)−1x0 is 1×1 and is therefore trace(BA). Note that x0 is p×1 and X is N×p, so that xT is approximately equal to σ2σ−2trace(Ip)/N = p/N. equal to its own trace. Therefore Ex0xT 0 (X T X)−1x0 = traceEx0(x0xT 0 (X T X)−1) which 8
分享到:
收藏