A. Answers to Exercises

Every exercise is answered here (at least briefly), and some of these answers go beyond what was asked. Readers will learn best if they make a serious attempt to find their own answers BEFORE PEEKING at this appendix.

The authors will be interested to learn of any solutions (or partial solutions) to the research problems, or of any simpler (or more correct) ways to solve the non-research ones.

(The first finder of every error in this book will receive a reward of $2.56.)

1.1 The proof is fine except when n = 2. If all sets of two horses have horses of the same color, the statement is true for any number of horses.

Does that mean I have to find every error?

1.2 If Xn is the number of moves, we have X0 = 0 and Xn = Xn1 + 1 + Xn1 + 1 + Xn1 when n > 0. It follows (for example by adding 1 to both sides) that Xn = 3n 1. (After Image moves, it turns out that the entire tower will be on the middle peg, halfway home!)

(We meant to say “any error.”)

Does that mean only one person gets a reward?

1.3 There are 3n possible arrangements, since each disk can be on any of the pegs. We must hit them all, since the shortest solution takes 3n 1 moves. (This construction is equivalent to a “ternary Gray code,” which runs through all numbers from (0 . . . 0)3 to (2 . . . 2)3, changing only one digit at a time.)

(Hmmm. Try it and see.)

1.4 No. If the largest disk doesn’t have to move, 2n1 1 moves will suffice (by induction); otherwise (2n1 1) + 1 + (2n1 1) will suffice (again by induction).

1.5 No; different circles can intersect in at most two points, so the fourth circle can increase the number of regions to at most 14. However, it is possible to do the job with ovals:

Image

The number of intersection points turns out to give the whole story; convexity was a red herring.

Venn [359] claimed that there is no way to do the five-set case with ellipses, but a five-set construction with ellipses was found by Grünbaum [167].

1.6 If the nth line intersects the previous lines in k > 0 distinct points, we get k 1 new bounded regions (assuming that none of the previous lines were mutually parallel) and two new infinite regions. Hence the maximum number of bounded regions is (n2)+(n3)+· · · = Sn2 = (n1)(n2)/2 = Ln2n.

This answer assumes that n > 0.

1.7 The basis is unproved; and in fact, H(1)2.

1.8 Q2 = (1 + β); Q3 = (1 + α + β)/αβ; Q4 = (1 + α); Q5 = α; Q6 = β. So the sequence is periodic!

1.9 (a) We get P(n 1) from the inequality

Image

(b) x1 . . . xnxn+1 . . . x2n (((x1 + · · · + xn)/n)((xn+1 + · · · + x2n)/n))n by P(n); the product inside is ((x1 + · · · +x2n)/2n)2 by P(2). (c) For example, P(5) follows from P(6) from P(3) from P(4) from P(2).

1.10 First show that Rn = Rn1 + 1 + Qn1 + 1 + Rn1, when n > 0. Incidentally, the methods of Chapter 7 will tell us that Qn = ((1 + Image)n+1 – (1Image)n+1)/(2Image) – 1.

1.11 (a) We cannot do better than to move a double (n 1)-tower, then move (and invert the order of) the two largest disks, then move the double (n 1)-tower again; hence An = 2An1 + 2 and An = 2Tn = 2n+1 2. This solution interchanges the two largest disks but returns the other 2n 2 to their original order.

(b) Let Bn be the minimum number of moves. Then B1 = 3, and it can be shown that no strategy does better than Bn = An1 +2+An1 +2+Bn1 when n > 1. Hence Bn = 2n+25, for all n > 0. Curiously this is just 2An1, and we also have Bn = An1 + 1 + An1 + 1 + An1 + 1 + An1.

1.12 If all mk > 0, then A(m1, . . . , mn) = 2A(m1, . . . , mn1)+mn. This is an equation of the “generalized Josephus” type, with solution (m1 . . . mn)2 = 2n1m1 + · · · + 2mn1 + mn.

Incidentally, the corresponding generalization of exercise 11b appears to satisfy the recurrence

Image

1.13 Given n straight lines that define Ln regions, we can replace them by extremely narrow zig-zags with segments sufficiently long that there are nine intersections between each pair of zig-zags. This shows that ZZn = ZZn1+9n8, for all n > 0; consequently ZZn = 9Sn8n + 1 = Imagen2Imagen + 1.

1.14 The number of new 3-dimensional regions defined by each new cut is the number of 2-dimensional regions defined in the new plane by its intersections with the previous planes. Hence Pn = Pn1 + Ln1, and it turns out that P5 = 26. (Six cuts in a cubical piece of cheese can make 27 cubelets, or up to P6 = 42 cuts of weirder shapes.)

Incidentally, the solution to this recurrence fits into a nice pattern if we express it in terms of binomial coefficients (see Chapter 5):

Image

I bet I know what happens in four dimensions!

Here Xn is the maximum number of 1-dimensional regions definable by n points on a line.

1.15 The function I satisfies the same recurrence as J when n > 1, but I(1) is undefined. Since I(2) = 2 and I(3) = 1, there’s no value of I(1) = α that will allow us to use our general method; the “end game” of unfolding depends on the two leading bits in n’s binary representation.

If n = 2m + 2m1 + k, where 0 k < 2m+1 + 2m (2m + 2m1) = 2m +2m1, the solution is I(n) = 2k+1 for all n > 2. Another way to express this, in terms of the representation n = 2m + l, is to say that

Image

1.16 Let g(n) = a(n)α + b(n)β0 + c(n)β1 + d(n)γ. We know from (1.18) that a(n)α + b(n)β0 + c(n)β1 = (α βbm1 βbm2 . . . βb1 βb0)3 when n = (1 bm1 . . . b1 b0)2; this defines a(n), b(n), and c(n). Setting g(n) = n in the recurrence implies that a(n) + c(n) d(n) = n; hence we know everything. [Setting g(n) = 1 gives the additional identity a(n) 2b(n) 2c(n) = 1, which can be used to define b(n) in terms of the simpler functions a(n) and a(n) + c(n).]

1.17 In general we have Wm 2Wmk + Tk, for 0 k m. (This relation corresponds to transferring the top m k, then using only three pegs to move the bottom k, then finishing with the top m k.) The stated relation turns out to be based on the unique value of k that minimizes the right-hand side of this general inequality, when m = n(n + 1)/2. (However, we cannot conclude that equality holds; many other strategies for transferring the tower are conceivable.) If we set Yn = (Wn(n+1)/2 1)/2n, we find that Yn Yn1 + 1; hence Wn(n+1)/2 2n(n 1) + 1.

1.18 It suffices to show that both of the lines from (n2j, 0) intersect both of the lines from (n2k, 0), and that all these intersection points are distinct.

A line from (xj, 0) through (xj aj, 1) intersects a line from (xk, 0) through (xk ak, 1) at the point (xj taj, t) where t = (xk xj)/(ak aj). Let xj = n2j and aj = nj + (0 or nn). Then the ratio t = (n2k n2j)/(nk nj + (–nn or 0 or nn)) lies strictly between nj + nk 1 and nj + nk + 1; hence the y coordinate of the intersection point uniquely identifies j and k. Also the four intersections that have the same j and k are distinct.

1.19 Not when n > 5. A bent line whose half-lines run at angles θ and θ + 30° from its apex can intersect four times with another whose half-lines run at angles ϕ and ϕ + 30° only if 30° < |θ ϕ| < 150°. We can’t choose more than 5 angles this far apart from each other. (It is possible to choose 5.)

1.20 Let h(n) = a(n)α + b(n)β0 + c(n)β1 + d(n)γ0 + e(n)γ1. We know from (1.18) that a(n)α + b(n)β0 + c(n)β1 = (α βbm1 βbm2 . . . βb1 βb0)4 when n = (1 bm1 . . . b1 b0)2; this defines a(n), b(n), and c(n). Setting h(n) = n in the recurrence implies that a(n) + c(n) 2d(n) 2e(n) = n; setting h(n) = n2 implies that a(n) + c(n) + 4e(n) = n2. Hence d(n) = (3a(n) + 3c(n) n2 2n)/4; e(n) = (n2 a(n) c(n)) /4.

1.21 We can let q be the least (or any) common multiple of 2n, 2n 1, . . . , n + 1. [A non-rigorous argument suggests that a “random” value of q will succeed with probability

Image

so we might expect to find such a q less than 4n.]

1.22 Take a regular polygon with 2n sides and label the sides with the elements of a “de Bruijn cycle” of length 2n. (This is a cyclic sequence of 0’s and 1’s in which all n-tuples of adjacent elements are different; see [207, exercise 2.3.4.2–23] and [208, exercise 3.2.2–17].) Attach a very thin convex extension to each side that’s labeled 1. The n sets are copies of the resulting polygon, rotated by the length of k sides for k = 0, 1, . . . , n 1.

I once rode a de Bruijn cycle (when visiting at his home in Nuenen, The Netherlands).

1.23 Yes. (We need principles of elementary number theory from Chapter 4.) Let L(n) = lcm(1, 2, . . . , n). We can assume that n > 2; hence by Bertrand’s postulate there is a prime p between n/2 and n. We can also assume that j > n/2, since q= L(n) + 1 q leaves j= n + 1 j if and only if q leaves j. Choose q so that q1 (mod L(n)/p) and qj + 1 n (mod p). The people are now removed in order 1, 2, . . . , n p, j + 1, j + 2, . . . , n, n p + 1, . . . , j 1.

1.24 The only known examples are: Xn = 2i sin πr + 1/Xn1, where r is rational and 0r < Image (all period lengths 2 occur as r varies); Gauss’s recurrence of period 5 in exercise 8; H. Todd’s even more remarkable recurrence Xn = (1+Xn1+Xn2)/Xn3, which has period 8 (see [261]); and recurrences derived from these when we replace Xn by a constant times Xmn. We can assume that the first nonzero coefficient in the denominator is unity, and that the first nonzero coefficient in the numerator (if any) has nonnegative real part. Computer algebra shows easily that there are no further solutions of period 5 when k = 2. A partial theory has been developed by Lyness [261, 262] and by Kurshan and Gopinath [231].

An interesting example of another type, with period 9 when the starting values are real, is the recurrence Xn = |Xn1|Xn2 discovered by Morton Brown [43]. Nonlinear recurrences having any desired period 5 can be based on continuants [65].

1.25 If T(k)(n) denotes the minimum number of moves needed to transfer n disks with k auxiliary pegs (hence T(1)(n) = Tn and T(2)(n) = Wn), we have Image. No examples (n, k) are known where this inequality fails to be an equality. When k is small compared with n, the formula 2n+1kImage gives a convenient upper bound on Image.

1.26 The execution-order permutation can be computed in O(n log n) steps for all q and n [209, exercises 5.1.1–2 and 5.1.1–5]. Bjorn Poonen has proved that non-Josephus sets with exactly four “bad guys” exist whenever n0 (mod 3) and n 9; in fact, the number of such sets is at least Image for some Image > 0. He also found by extensive computations that the only other n < 24 with non-Josephus sets is n = 20, which has 236 such sets with k = 14 and two with k = 13. (One of the latter is {1, 2, 3, 4, 5, 6, 7, 8, 11, 14, 15, 16, 17}; the other is its reflection with respect to 21.) There is a unique non-Josephus set with n = 15 and k = 9, namely {3, 4, 5, 6, 8, 10, 11, 12, 13}.

2.1 There’s no agreement about this; three answers are defensible: (1) We can say that Image is always equivalent to mkn qk; then the stated sum is zero. (2) A person might say that the given sum is q4 + q3 + q2 + q1 + q0, by summing over decreasing values of k. But this conflicts with the generally accepted convention that Image when n = 0. (3) We can say that Image; then the stated sum is equal to q1 q2 q3. This convention may appear strange, but it obeys the useful law Image for all a, b, c.

It’s best to use the notation Image only when n m 1; then both conventions (1) and (3) agree.

2.2 This is |x|. Incidentally, the quantity ([x > 0] [x < 0]) is often called sign(x) or signum(x); it is +1 when x > 0, 0 when x = 0, and 1 when x < 0.

2.3 The first sum is, of course, a0 + a1 + a2 + a3 + a4 + a5; the second is a4 + a1 + a0 + a1 + a4, because the sum is over the values k {2, 1, 0, +1, +2}. The commutative law doesn’t apply here because the function p(k) = k2 is not a permutation. Some values of n (e.g., n = 3) have no k such that p(k) = n; others (e.g., n = 4) have two such k.

2.4 (a) Image

(b) Image

2.5 The same index ‘k’ is being used for two different index variables, although k is bound in the inner sum. This is a famous mistake in mathematics (and computer programming). The result turns out to be correct if aj = ak for all j and k, 1 j, k n.

2.6 It’s [1 j n](n j + 1). The first factor is necessary here because we should get zero when j < 1 or j > n.

2.7 Image. A version of finite calculus based on instead of Δ would therefore give special prominence to rising factorial powers.

2.8 0, if m 1; 1/|m|!, if m 0.

2.9 Image, for integers m and n. Setting m = n tells us that Image.

2.10 Another possible right-hand side is Eu Δv + v Δu.

2.11 Break the left-hand side into two sums, and change k to k + 1 in the second of these.

2.12 If p(k) = n then n + c = k + ((–1)k + 1)c and ((–1)k + 1) is even; hence (–1)n+c = (–1)k and k = n (–1)n+cc. Conversely, this value of k yields p(k) = n.

2.13 Let R0 = α, and Rn = Rn1 + (–1)n(β + + n2δ) for n > 0. Then R(n) = A(n)α + B(n)β + C(n)γ + D(n)δ. Setting Rn = 1 yields A(n) = 1. Setting Rn = (–1)n yields A(n) + 2B(n) = (–1)n. Setting Rn = (–1)nn yields B(n)+2C(n) = (–1)nn. Setting Rn = (–1)nn2 yields B(n)–2C(n) + 2D(n) = (–1)nn2. Therefore 2D(n) = (–1)n(n2+n); the stated sum is D(n).

2.14 The suggested rewrite is legitimate since we have k = 1jk 1 when 1 k n. Sum first on k; the multiple sum reduces to

Image

2.15 The first step replaces k(k + 1) by 21jk j. The second step gives Image.

2.16 Image, by (2.52).

2.17 Use induction for the first two =’s, and (2.52) for the third. The second line follows from the first.

2.18 Use the facts that (z)+ |z|, (z) |z|, (z)+ |z|, (z) |z|, and |z| (z)+ + (z) + (z)+ + (z).

2.19 Multiply both sides by 2n1/n! and let Sn = 2nTn/n! = Sn1 + 3 · 2n1 = 3(2n 1) + S0. The solution is Tn = 3 · n! + n!/2n1. (We’ll see in Chapter 4 that Tn is an integer only when n is 0 or a power of 2.)

“It is a profoundly erroneous truism, repeated by all copybooks and by eminent people when they are making speeches, that we should cultivate the habit of thinking of what we are doing. The precise opposite is the case. Civilization advances by extending the number of important operations which we can perform without thinking about them. Operations of thought are like cavalry charges in a battle—they are strictly limited in number, they require fresh horses, and must only be made at decisive moments.”

—A. N. Whitehead [370]

2.20 The perturbation method gives

Image

2.21 Extracting the final term of Sn+1 gives Sn+1 = 1 Sn; extracting the first term gives

Image

Hence 2Sn = 1 + (–1)n and we have Sn = [n is even]. Similarly, we find

Image

hence 2Tn = n + 1 Sn and we have Tn = Image(n + [n is odd]). Finally, the same approach yields

Un+1 = (n + 1)2Un = Un + 2Tn + Sn
  = Un + n + [n is odd] + [n is even]
  = Un + n + 1.

Hence Un is the triangular number Image.

2.22 Twice the general sum gives a “vanilla” sum over 1 j, k n, which splits and yields twice (k akAk)(k bkBk) (k akBk) (k bkAk).

2.23 (a) This approach gives four sums that evaluate to 2n + Hn – 2n + (Hn + Image1). (It would have been easier to replace the summand by 1/k + 1/(k + 1).) (b) Let u(x) = 2x + 1 and Δv(x) = 1/x(x + 1) = (x 1)2; then Δu(x) = 2 and v(x) = –(x 1)1 = –1/x. The answer is Image.

2.24 Summing by parts, ΣxmHx δx = xm + 1Hx/(m+1)–xm+1/(m+1)2 +C; hence Σ0k<n kmHk = nm + 1 (Hn 1/(m + 1))/(m + 1) + 0m + 1/(m + 1)2. In our case m = 2, so the sum comes to 1 (Hn + 1)/(n + 1).

2.25 Here are some of the basic analogies:

Image

2.26 P2 = (1j,kn ajak)(1j=kn ajak). The first factor is equal to Image; the second factor is Image. Hence Image.

2.27 Δ(cx) = cx(c x 1) = cx+2/(c x). Setting c = 2 and decreasing x by 2 yields Δ(–(–2)x2) = (–2)x/x, hence the stated sum is (–2)1 – (–2)n1 = (–1)nn! 1.

2.28 The interchange of summation between the second and third lines is not justifiable; the terms of this sum do not converge absolutely. Everything else is perfectly correct, except that the result of k1[k = j 1]k/j should perhaps have been written [j 1 1](j 1)/j and simplified explicitly.

As opposed to imperfectly correct.

2.29 Use partial fractions to get

Image

The (–1)k factor now makes the two halves of each term cancel with their neighbors. Hence the answer is 1/4 + (–1)n/(8n + 4).

2.30 Image. So we have

(b a)(b + a 1) = 2100 = 22 ·3·52 ·7.

There is one solution for each way to write 2100 = x · y where x is even and y is odd; we let Image and Image. So the number of solutions is the number of divisors of 3 · 52 · 7, namely 12. In general, there are p>2(np + 1) ways to represent p pnp, where the products range over primes.

2.31 j,k2 jk = j2 1/j2(1 1/j) = j2 1/j(j 1). The second sum is, similarly, 3/4.

2.32 If 2n x < 2n+1, the sums are 0+· · ·+n+(xn1)+· · ·+(x2n) = n(xn) = (x1) + (x3) + · · · + (x2n+1). If 2n 1 x < 2n they are, similarly, both equal to n(x n). (Looking ahead to Chapter 3, the formula ImageImage(x + 1)Image(xImageImage(x + 1Image) covers both cases.)

2.33 If K is empty, kK ak = . The basic analogies are:

Image

A permutation that consumes terms of one sign faster than those of the other can steer the sum toward any value that it likes.

2.34 Let K+ = {k | ak 0} and K = {k | ak < 0}. Then if, for example, n is odd, we choose Fn to be Fn1En, where EnK is sufficiently large that k(Fn1K+) ak kEn (–ak) < A.

2.35 Goldbach’s sum can be shown to equal

Image

as follows: By unsumming a geometric series, it equals kP, l1 kl; therefore the proof will be complete if we can find a one-to-one correspondence between ordered pairs (m, n) with m, n 2 and ordered pairs (k, l) with kP and l 1, where mn = kl when the pairs correspond. If mP we let (m, n) (mn, 1); but if m = abP, we let (m, n) (an, b).

With this self-description, Golomb’s sequence wouldn’t do too well on the Dating Game.

2.36 (a) By definition, g(n) g(n 1) = f(n). (b) By part (a), g(g(n)) g(g(n 1)) = k f(k)[g(n 1) < k g(n)] = n(g(n) g(n 1)) = nf(n). (c) By part (a) again, g(g(g(n)))g(g(g(n 1))) is

Image

Colin Mallows observes that the sequence can also be defined by the recurrence

f(1) = 1;     f(n + 1) = 1 + f (n + 1f(f(n))),   for n0.

2.37 (RLG thinks they probably won’t fit; DEK thinks they probably will; OP is not committing himself.)

3.1 m = Imagelg nImage; l = n 2m = n 2Imagelg nImage.

3.2 (a) Imagex + .5Image. (b) Imagex .5Image.

3.3 This is Imagemn {}n/αImage = mn 1, since 0 < {} < 1.

3.4 Something where no proof is required, only a lucky guess (I guess).

3.5 We have ImagenxImage = ImagenImagexImage + n{x}Image = nImagexImage + Imagen{x}Image by (3.8) and (3.6). Therefore ImagenxImage = nImagexImage Imagen{x}Image = 0 0 n{x} < 1 {x} < 1/n, assuming that n is a positive integer. (Notice that nImagexImage ImagenxImage for all x in this case.)

3.6 Imagef(x)Image = Imagef(ImagexImage)Image.

3.7 Imagen/mImage + n mod m.

3.8 If all boxes contain < Imagen/mImage objects, then n (Imagen/mImage 1) m, so n/m + 1 Imagen/mImage, contradicting (3.5). The other proof is similar.

3.9 We have m/n1/q = (n mumble m)/qn. The process must terminate, because 0 n mumble m < m. The denominators of the representation are strictly increasing, hence distinct, because qn/(n mumble m) > q.

3.10 Imagex + ImageImage – [(2x + 1)/4 is not an integer] is the nearest integer to x, if {x} ≠ Image; otherwise it’s the nearest even integer. (See exercise 2.) Thus the formula gives an “unbiased” way to round.

3.11 If n is an integer, α < n < β ImageαImage < n < ImageβImage. The number of integers satisfying a < n < b when a and b are integers is (b a 1)[b > a]. We would therefore get the wrong answer if α = β = integer.

3.12 Subtract Imagen/mImage from both sides, by (3.6), getting Image(n mod m)/mImage = Image(n mod m + m 1)/mImage. Both sides are now equal to [n mod m > 0], since 0 n mod m < m.

A shorter but less direct proof simply observes that the first term in (3.24) must equal the last term in (3.25).

3.13 If they form a partition, the text’s formula for N(α, n) implies that 1/α + 1/β = 1, because the coefficients of n in the equation N(α, n) + N(β, n) = n must agree if the equation is to hold for large n. Hence α and β are both rational or both irrational. If both are irrational, we do get a partition, as shown in the text. If both can be written with numerator m, the value m 1 occurs in neither spectrum, and m occurs in both. (However, Golomb [151] has observed that the sets {ImageImage | n 1} and {ImageImage 1 | n 1} always do form a partition, when 1/α + 1/β = 1.)

3.14 It’s obvious by (3.22) if ny = 0, otherwise true by (3.21) and (3.6).

3.15 Plug in ImagemxImage for n in (3.24): ImagemxImage = ImagexImage + ImagexImageImage + ··· + ImagexImageImage.

3.16 The formula n mod 3 = 1 + Image((ω – 1n – (ω + 22n) can be verified by checking it when 0 n < 3.

A general formula for n mod m, when m is any positive integer, appears in exercise 7.25.

3.17 Σj,k[0 k < m][1 j x + k/m] = Σj,k[0 k< m][1 j ImagexImage] × [k m(j x)] = Σ1jImagexImage Σk[0 k < m] Σj=ImagexImage Σk [0 k < m(j x)] = mImagexImage Imagem(ImagexImage x)Image = ImagemxImage = ImagemxImage.

3.18 We have

Image

If j 1 v, there is no contribution, because (j + v)α1 n. Hence j = ImageImage is the only case that matters, and the value in that case equals Image(ImageImage + v)α1Image n Image1Image.

3.19 If and only if b is an integer. (If b is an integer, logb x is a continuous, increasing function that takes integer values only at integer points. If b is not an integer, the condition fails when x = b.)

3.20 We have k kx[α kx β] = x k k[Imageα/xImage k Imageβ/xImage], which sums to Imagex(Imageβ/xImage Imageβ/x + 1Image Imageα/xImage Imageα/x 1Image).

3.21 If 10n 2M < 10n+1, there are exactly n+1 such powers of 2, because there’s exactly one such k-digit power of 2 for each k. Therefore the answer is 1 + ImageM log 2Image.

Note: The number of powers of 2 with leading digit l is more difficult, when l > 1; it’s 0nM (Imagen log 2 log lImage Imagen log 2 log(l + 1)Image).

3.22 All terms are the same for n and n1 except the kth, where n = 2k–1q and q is odd; we have Sn = Sn1 + 1 and Tn = Tn1 + 2kq. Hence Sn = n and Tn = n(n + 1).

3.23 Xn = mImagem(m1) < nImagem(m + 1) ⇔ m2m + Image < 2n < m2 + m + ImagemImage < Image < m + Image.

3.24 Let β = α/(α + 1). Then the number of times the nonnegative integer m occurs in Spec(β) is exactly one more than the number of times it occurs in Spec(α). Why? Because N(β, n) = N(α, n) + n + 1.

3.25 Continuing the development in the text, if we could find a value of m such that Km m, we could violate the stated inequality at n + 1 when n = 2m + 1. (Also when n = 3m + 1 and n = 3m + 2.) But the existence of such an m = n+ 1 requires that 2KImagen/2Image n′ or 3KImagen/3Image n′, i.e., that

KImagen′/2ImageImagen′/2Image    or    KImagen′/3ImageImagen′/3Image.

Aha. This goes down further and further, implying that K0 0; but K0 = 1.

What we really want to prove is that Kn is strictly greater than n, for all n 0. In fact, it’s easy to prove this by induction, although it’s a stronger result than the one we couldn’t prove!

(This exercise teaches an important lesson. It’s more an exercise about the nature of induction than about properties of the floor function.)

“In trying to devise a proof by mathematical induction, you may fail for two opposite reasons. You may fail because you try to prove too much: Your P(n) is too heavy a burden. Yet you may also fail because you try to prove too little: Your P(n) is too weak a support. In general, you have to balance the statement of your theorem so that the support is just enough for the burden.”

—G. Pólya [297]

3.26 Induction, using the stronger hypothesis

Image

3.27 If Image = 2mba, where a is 0 or 1, then Image = 3mba.

3.28 The key observation is that an = m2 implies an+2k+1 = (m + k)2 + mk and an+2k+2 = (m+k)2+2m, for 0 k m; hence an+2m+1 = (2m)2. The solution can be written in a nice form discovered by Carl Witty:

Image

3.29 D(α, ImageαnImage) is at most the maximum of the absolute value of s(α, ImageImage, v′) = s(α, n, v) S + Image + {0 or 1} + v {0 or 1}.

3.30 Xn = α2n + α2n, by induction; and Xn is an integer.

This logic is seriously floored.

3.31 Here’s an “elegant,” “impressive” proof that gives no clue about how it was discovered:

ImagexImage + ImageyImage + Imagex + yImage = Imagex + ImageyImageImage + Imagex + yImage
  Imagex + ImageImage2yImageImage + Imagex + ImageImage2yImage + ImageImage
  = Image2x + Image2yImageImage = Image2xImage + Image2yImage.

But there’s also a simple, graphical proof based on the observation that we need to consider only the case 0 x, y < 1. Then the functions look like this in the plane:

Image

A slightly stronger result is possible, namely

ImagexImage + ImageyImage + Imagex + yImage Image2xImage + Image2yImage;

but this is stronger only when Image. If we replace (x, y) by (–x, x + y) in this identity and apply the reflective law (3.4), we get

ImageyImage + Imagex + yImage + Image2xImage ImagexImage + Image2x + 2yImage.

3.32 Let f(x) be the sum in question. Since f(x) = f(–x), we may assume that x 0. The terms are bounded by 2k as k and by x2/2k as k +, so the sum exists for all real x.

We have f(2x) = 2 k 2k1x/2k12 = 2f(x). Let f(x) = l(x) + r(x) where l(x) is the sum for k 0 and r(x) is the sum for k > 0. Then l(x + 1) = l(x), and l(x) 1/2 for all x. When 0 x < 1, we have r(x) = x2/2 + x2/4 + · · · = x2, and r(x + 1) = (x 1)2/2 + (x + 1)2/4 + (x + 1)2/8 + · · · = x2 + 1. Hence f(x + 1) = f(x) + 1, when 0 x < 1.

We can now prove by induction that f(x + n) = f(x) + n for all integers n 0, when 0 x < 1. In particular, f(n) = n. Therefore in general, f(x) = 2mf(2mx) = 2mImage2mxImage + 2mf ({2mx}). But f({2mx}) = l({2mx}) + r({2mx}) ≤ Image + 1; so |f(x) – x| ≤ |2m Image2mxImagex| + 2m·Image ≤ 2m · Image for all integers m.

The inescapable conclusion is that f(x) = |x| for all real x.

3.33 Let r = nImage be the radius of the circle. (a) There are 2n1 horizontal lines and 2n1 vertical lines between cells of the board, and the circle crosses each of these lines twice. Since r2 is not an integer, the Pythagorean theorem tells us that the circle doesn’t pass through the corner of any cell. Hence the circle passes through as many cells as there are crossing points, namely 8n 4 = 8r. (The same formula gives the number of cells at the edge of the board.) (b) f(n,k) = 4ImageImageImage.

It follows from (a) and (b) that

Image

The task of obtaining more precise estimates of this sum is a famous problem in number theory, investigated by Gauss and many others; see Dickson [78, volume 2, chapter 6].

3.34 (a) Let m = Imagelg nImage. We can add 2m n terms to simplify the calculations at the boundary:

Image

Consequently f(n) = nm 2m + 1.

(b) We have Imagen/2Image = Image(n+1)/2Image, and it follows that the solution to the general recurrence g(n) = a(n) + g (Imagen/2Image) + g(Imagen/2Image) must satisfy Δg(n) = Δa(n)+Δg (Imagen/2Image). In particular, when a(n) = n1, Δf(n) = 1+Δf (Imagen/2Image) is satisfied by the number of bits in the binary representation of n, namely Imagelg(n + 1)Image. Now convert from Δ to Σ.

A more direct solution can be based on the identities Imagelg 2jImage = Imagelg jImage + 1 and Imagelg(2j 1)Image = Imagelg jImage + [j > 1], for j 1.

3.35 (n + 1)2n!e = An + (n + 1)2 + (n + 1) + Bn, where

Image

is a multiple of n and

Image

is less than 1. Hence the answer is 2 mod n.

3.36 The sum is

Image

3.37 First consider the case m < n, which breaks into subcases based on whether Image; then show that both sides change in the same way when m is increased by n.

This is really only a level 4 problem, in spite of the way it’s stated.

3.38 At most one xk can be noninteger. Discard all integer xk, and suppose that n are left. When {x}0, the average of {mx} as m lies between Image and Image; hence {mx1} + · · · + {mxn} {mx1 + · · · + mxn} cannot have average value zero when n > 1.

But the argument just given relies on a difficult theorem about uniform distribution. An elementary proof is possible, sketched here for n = 2: Let Pm be the point ({mx}, {my}). Divide the unit square 0 x, y < 1 into triangular regions A and B according as x + y < 1 or x + y 1. We want to show that PmB for some m, if {x} and {y} are nonzero. If P1B, we’re done. Otherwise there is a disk D of radius Image > 0 centered at P1 such that DA. By Dirichlet’s box principle, the sequence P1, . . . , PN must contain two points with |Pk Pj| < Image and k > j, if N is large enough.

Image

It follows that Pkj1 is within Image of (1, 1) P1; hence Pkj1B.

3.39 Replace j by b j and add the term j = 0 to the sum, so that exercise 15 can be used for the sum on j. The result,

Imagex/bkImage Imagex/bk+1Image + b1,

telescopes when summed on k.

3.40 Let Image2ImageImage = 4k + r where 2 r < 2, and let m = ImageImageImage. Then the following relationships can be proved by induction:

Image

Thus, when k 1, Wk is a segment of length 2k 1 where the path travels west and y(n) = k; Sk is the interior of a segment of length 2k where the path travels south and x(n) = k; etc. (a) The desired formula is therefore

y(n) = (–1)m((nm(m + 1))·[Image2ImageImage is odd] – ImageImagemImage).

(b) On all segments, k = max(|x(n)|, |y(n)|). On segments Wk and Sk we have x < y and n + x + y = m(m + 1) = (2k)2 2k; on segments Ek and Nk we have x y and n x y = m(m + 1) = (2k)2 + 2k. Hence the sign is (–1)[x(n)<y(n)].

3.41 Since 1/ϕ + 1/ϕ2 = 1, the stated sequences do partition the positive integers. Since the condition g(n) = f(f(n)) + 1 determines f and g uniquely, we need only show that ImageImageImageϕImage + 1 = Image2Image for all n > 0. This follows from exercise 3, with α = ϕ and n = 1.

3.42 No; an argument like the analysis of the two-spectrum case in the text and in exercise 13 shows that a tripartition occurs if and only if 1/α + 1/β + 1/γ = 1 and

Image

for all n > 0. But the average value of {(n + 1)} is 1/2 if α is irrational, by the theorem on uniform distribution. The parameters can’t all be rational, and if γ = m/n the average is 3/2 1/(2n). Hence γ must be an integer, but this doesn’t work either. (There’s also a proof of impossibility that uses only simple principles, without the theorem on uniform distribution; see [155].)

3.43 One step of unfolding the recurrence for Kn gives the minimum of the four numbers 1 + a + a · b · KImage(n1a)/(a·b)Image, where a and b are each 2 or 3. (This simplification involves an application of (3.11) to remove floors within floors, together with the identity x + min(y, z) = min(x + y, x + z). We must omit terms with negative subscripts; i.e., with n 1 a < 0.)

Continuing along such lines now leads to the following interpretation: Kn is the least number > n in the multiset S of all numbers of the form

1 + a1 + a1a2 + a1a2a3 + · · · + a1a2a3 . . . am ,

where m 0 and each ak is 2 or 3. Thus,

S = {1, 3, 4, 7, 9, 10, 13, 15, 19, 21, 22, 27, 28, 31, 31, . . . };

the number 31 is in S “twice” because it has two representations 1 + 2 + 4 + 8 + 16 = 1 + 3 + 9 + 18. (Incidentally, Michael Fredman [134] has shown that limn→∞ Kn/n = 1, i.e., that S has no enormous gaps.)

3.44 Let Image mumble(q1), so that Image and Image. Now Image, and the results follow. (This is the solution found by Euler [116], who determined the a’s and d’s sequentially without realizing that a single sequence Image would suffice.)

Too easy.

3.45 Let α > 1 satisfy α + 1/α = 2m. Then we find 2Yn = α2n + α2n, and it follows that Yn = Imageα2n/2Image.

3.46 The hint follows from (3.9), since 2n(n + 1) = Image2(n + Image)2Image. Let n + θ = (Imagel + Imagel–l)m and n′ + θ′ = (Imagel+1 + Imagel)m, where 0 θ, θ′ < 1. Then θ= mod 1 = d, where d is 0 or 1. We want to prove that n′ = ImageImage(n + Image)Image; this equality holds if and only if

0 ≤ θ′(2Image) + Image(1 – d) < 2.

To solve the recurrence, note that Spec(1 + 1/Image) and Spec(1 + Image) partition the positive integers; hence any positive integer a can be written uniquely in the form a = Image(Imagel + Imagel–1)mImage, where l and m are integers with m odd and l0. It follows that Ln = Image(Imagel+n + Imagel+n–1)mImage.

3.47 (a) Image. (b) c is an integer. (c) c = 0. (d) c is arbitrary. See the answer to exercise 1.2.4–40 in [207] for more general results.

3.48 Let x:0 = 1 and x:(k+1) = xImagex:kImage; also let ak = {x:k} and bk = Imagex:kImage, so that the stated identity reads x3 = 3x:3 + 3a1a2 + Image3b1b2 + Image. Since ak + bk = x:k = xbk1 for k 0, we have (1 xz)(1 + b1z + b2z2 + · · ·) = 1 a1z a2z2 · · · ; thus

Image

Take the logarithm of both sides, to separate the a’s from the b’s. Then differentiate with respect to z, obtaining

Image

The coefficient of zn1 on the left is xn; on the right it is a formula that matches the given identity when n = 3.

Similar identities for the more general product x0x1 . . . xn1 can also be derived [170].

3.49 (Solution by Heinrich Rolletschek.) We can replace (α, β) by ({β}, α + ImageβImage) without changing ImageImage + ImageImage. Hence the condition α = {β} is necessary. It is also sufficient: Let m = ImageβImage be the least element of the given multiset, and let S be the multiset obtained from the given one by subtracting mn from the nth smallest element, for all n. If α = {β}, consecutive elements of S differ by either 0 or 2, hence the multiset ImageS = Spec(α) determines α.

A more interesting (still unsolved) problem: Restrict both α and β to be < 1, and ask when the given multiset determines the unordered pair {α, β}.

3.50 According to unpublished notes of William A. Veech, it is sufficient to have αβ, β, and 1 linearly independent over the rationals.

3.51 H. S. Wilf observes that the functional equation f(x21) = f(x)2 would determine f(x) for all x ϕ if we knew f(x) on any interval (ϕ . . ϕ + Image).

3.52 There are infinitely many ways to partition the positive integers into three or more generalized spectra with irrational αk; for example,

Spec(; 0) ∪ Spec(; α) ∪ Spec(; ) ∪ Spec(β; 0)

works. But there’s a precise sense in which all such partitions arise by “expanding” a basic one, Spec(α) ∪ Spec(β); see [158]. The only known rational examples, e.g.,

Spec(7; – 3) ∪ Spec(Image; – 1) ∪ Spec(Image; 0) ,

are based on parameters like those in the stated conjecture, which is due to A. S. Fraenkel [128].

3.53 Partial results are discussed in [95, pages 30–31]. The greedy algorithm probably does not terminate.

4.1 1, 2, 4, 6, 16, 12.

4.2 Note that mp + np = min(mp, np) + max(mp, np). The recurrence lcm(m, n) = (n/(n mod m)) lcm(n mod m, m) is valid but not really advisable for computing lcm’s; the best way known to compute lcm(m, n) is to compute gcd(m, n) first and then to divide mn by the gcd.

4.3 This holds if x is an integer, but π(x) is defined for all real x. The correct formula,

π(x) π(x 1) = [ImagexImage is prime] ,

is easy to verify.

4.4 Between Image and Image we’d have a left-right reflected Stern–Brocot tree with all denominators negated, etc. So the result is all fractions m/n with mn. The condition mnmn= 1 still holds throughout the construction. (This is called the Stern–Brocot wreath, because we can conveniently regard the final Image as identical to the first Image, thereby joining the trees in a cycle at the top. The Stern–Brocot wreath has interesting applications to computer graphics because it represents all rational directions in the plane.)

4.5 Image and Image; this holds even when k < 0. (We will find a general formula for any product of L’s and R’s in Chapter 6.)

After all, ‘mod y’ sort of means “pretend y is zero.” So if it already is, there’s nothing to pretend.

4.6 a = b. (Chapter 3 defined x mod 0 = x, primarily so that this would be true.)

4.7 We need m mod 10 = 0, m mod 9 = k, and m mod 8 = 1. But m can’t be both even and odd.

4.8 We want 10x + 6y10x + y (mod 15); hence 5y0 (mod 15); hence y0 (mod 3). We must have y = 0 or 3, and x = 0 or 1.

4.9 32k+1 mod 4 = 3, so (32k+1 1)/2 is odd. The stated number is divisible by (37 1)/2 and (311 1)/2 (and by other numbers).

4.10 Image.

4.11 σ(0) = 1; σ(1) = 1; σ(n) = 0 for n > 1. (Generalized Möbius functions defined on arbitrary partially ordered structures have interesting and important properties, first explored by Weisner [366] and developed by many other people, notably Gian-Carlo Rota [313].)

4.12 dm kd μ(d/k) g(k) = km d(m/k) μ(d) g(k) = kmg(k)×[m/k = 1] = g(m), by (4.7) and (4.9).

4.13 (a) np 1 for all p; (b) μ(n)0.

4.14 True when k > 0. Use (4.12), (4.14), and (4.15).

4.15 No. For example, en mod 5 = [2 or 3]; en mod 11 = [2, 3, 7, or 10].

4.16 1/e1 + 1/e2 + · · · + 1/en = 1 1/(en(en 1)) = 1 1/(en+1 1).

4.17 We have fn mod fm = 2; hence gcd(fn, fm) = gcd(2, fm) = 1. (Incidentally, the relation fn = f0f1 . . . fn1 + 2 is very similar to the recurrence that defines the Euclid numbers en.)

4.18 If n = qm and q is odd, 2n+1 = (2m+1)(2nm 2n2m +· · ·2m+1).

4.19 The first sum is π(n), since the summand is [k + 1 is prime]. The inner sum in the second is 1k<m[km], so it is greater than 1 if and only if m is composite; again we get π(n). Finally Image, so the third sum is an application of Wilson’s theorem. To evaluate π(n) by any of these formulas is, of course, sheer lunacy.

4.20 Let p1 = 2 and let pn be the smallest prime greater than 2pn1. Then 2pn1 < pn < 2pn1 +1, and it follows that we can take b = limn→∞ lg(n) pn where lg(n) is the function lg iterated n times. The stated numerical value comes from p2 = 5, p3 = 37. It turns out that p4 = 237 + 9, and this gives the more precise value

b1.2516475977905

(but no clue about p5).

4.21 By Bertrand’s postulate, Pn < 10n. Let

Image

Then 10n2KPn + fraction (mod 102n1).

4.22 (bmn 1)/(b 1) = ((bm 1)/(b 1)) (bmnm + · · · + 1). [The only prime numbers of the form (10p 1)/9 for p < 49081 occur when p = 2, 19, 23, 317, 1031.] Numbers of this form are called “repunits.”

4.23 ρ(2k + 1) = 0; ρ(2k) = ρ(k) + 1, for k 1. By induction we can show that ρ(n) = ρ(n 2m), if n > 2m and m > ρ(n). The kth Hanoi move is disk ρ(k), if we number the disks 0, 1, . . . , n 1. This is clear if k is a power of 2. And if 2m < k < 2m+1, we have ρ(k) < m; moves k and k 2m correspond in the sequence that transfers m + 1 disks in Tm + 1 + Tm steps.

4.24 The digit that contributes dpm to n contributes dpm1 + · · · + d = d(pm 1)/(p 1) to Imagep(n!), hence Imagep(n!) = (n νp(n))/(p 1).

4.25 m\n mp = 0 or mp = np, for all p. It follows that (a) is true. But (b) fails, in our favorite example m = 12, n = 18. (This is a common fallacy.)

4.26 Yes, since Image defines a subtree of the Stern–Brocot tree.

4.27 Extend the shorter string with M’s (since M lies alphabetically between L and R) until both strings are the same length, then use dictionary order. For example, the topmost levels of the tree are LL < LM < LR < MM < RL < RM < RR. (Another solution is to append the infinite string RL to both inputs, and to keep comparing until finding L < R.)

4.28 We need to use only the first part of the representation:

Image

The fraction Image appears because it’s a better upper bound than Image, not because it’s closer than Image. Similarly, Image is a better lower bound than Image. The simplest upper bounds and the simplest lower bounds all appear, but the next really good approximation doesn’t occur until just before the string of R’s switches back to L.

4.29 1/α. To get 1 x from x in binary notation, we interchange 0 and 1; to get 1/α from α in Stern–Brocot notation, we interchange L and R. (The finite cases must also be considered, but they must work since the correspondence is order preserving.)

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.145.74.54