Skip to main content
Mathematics LibreTexts

2.6: Exercises

  • Page ID
    60305
  • \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \) \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)\(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\) \(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\)\(\newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    Exercise \(\PageIndex{1}\)

    Apply the division algorithm to the following number pairs. (Hint: replace negative numbers by positive ones.)

    1. \(110, 7\).
    2. \(51, -30\).
    3. \(-138, 24\).
    4. \(272, 119\).
    5. \(2378, 1769\).
    6. \(270, 175560\).

    Exercise \(\PageIndex{2}\)

    In this exercise we will exhibit the division algorithm applied to polynomials \(x+1\) and \(3x^3+2x+1\) with coefficients in \(\mathbb{Q}, \mathbb{R}\), or \(\mathbb{C}\).

    1. Apply long division to divide \(3021\) by \(11\). (Hint: \(3021 = 11 \cdot 275 -4\).)
    2. Apply the exact same algorithm to divide \(3x^3+2x+1\) by \(x+1\). In this algorithm, \(x^k\) behaves as \(10^k\) in (a). (Hint: every step, cancel the highest power of \(x\).)
    3. Verify that you obtain \(3x^3+2x+1 = (x+1)(3x^2-3x+5)-4\).
    4. Show that in general, if \(p_{1}\) and \(p_{2}\) are polynomials such that the degree of \(p_{1}\) is greater or equal to the degree of \(p_{2}\), then \(p_{1} = q_{2}p_{2}+p_{3}\), where the degree of \(p_{3}\) is less than the degree of \(p_{2}\). (Hint: perform long division as in (b). Stop when the degree of the remainder is less than that of \(p_{2}\).)
    5. Why does this division not work for polynomials with coefficients in \(\mathbb{Z}\)? (Hint: replace \(x+1\) by \(2x+1\).)
    1. Exercise \(\PageIndex{3}\)
    2. Compute by long division that \(3021 = 11 \cdot 274+7\).
    3. Conclude from exercise 2.2 that \(3021 = 11(300-30+5)-4\). (Hint: let \(x = 10\).)
    4. Conclude from exercise 2.2 that \(3 \cdot 163 +2 \cdot 16+1 = 17(3 \cdot 162-3 \cdot 16+5)-4\). (Hint: let \(x = 16\).)

    Exercise \(\PageIndex{4}\)

    1. Use unique factorization to show that any composite number \(n\) must have a prime factor less than or equal to \(\sqrt{n}\).
    2. Use that fact to prove: If we apply Eratosthenes’ sieve to \(\{2, 3, \cdots, n\}\), it is sufficient to sieve out numbers less than or equal to \(\sqrt{n}\).

    Exercise \(\PageIndex{5}\)

    We give an elementary proof of Corollary 2.15.

    1. Show that \(a \cdot \frac{b}{\gcd (a,b)}\) is a multiple of \(a\).
    2. Show that \(\frac{b}{\gcd (a,b)} \cdot b\) is a multiple of \(b\).
    3. Conclude that \(\frac{ab}{\gcd (a,b)}\) is a multiple of both \(a\) and \(b\) and thus greater than or equal to \(\mbox{lcm} (a,b)\).
    4. Show that \(a/(\frac{ab}{\mbox{lcm} (a,b)}) = \frac{\mbox{lcm} (a,b)}{b}\) is an integer. Thus \(\frac{ab}{\mbox{lcm} (a,b)}\) is a divisor of \(a\).
    5. Similarly, show that \(\frac{ab}{\mbox{lcm} (a,b)}\) is a divisor of \(b\).
    6. Conclude that \(\frac{ab}{\mbox{lcm} (a,b)} \le \gcd (a,b)\).
    7. Finish the proof.

    Exercise \(\PageIndex{6}\)

    It is possible to extend the definition of \(\gcd\) and \(\mbox{lcm}\) to more than two integers (not all of which are zero). For example \(\gcd (24, 27, 54) = 3\).

    1. Compute \(\gcd (6,10,15)\) and \(\mbox{lcm} (6,10,15)\).
    2. Give an example of a triple whose \(\gcd\) is one, but every pair of which has a \(\gcd\) greater than one.
    3. Show that there is no triple \(\{a,b,c\}\) whose \(\mbox{lcm}\) equals \(abc\), but every pair of which has lcm less than the product of that pair. (Hint: consider \(\mbox{lcm} (a \cdot b) \cdot c\).)

    Exercise \(\PageIndex{7}\)

    1. Give the prime factorization of the following numbers: \(12, 392, 1043, 31, 128, 2160, 487\).
    2. Give the prime factorization of the following numbers: \(12 \cdot 392, 1043 \cdot 31, 128 \cdot 2160\).
    3. Give the prime factorization of: \(1, 250000, 63^3, 720\), and the product of the last three numbers.

    Exercise \(\PageIndex{8}\)

    Use the Fundamental Theorem of Arithmetic to prove:

    1. Be ́zout’s Lemma.
    2. Euclid’s Lemma.

    Exercise \(\PageIndex{9}\)

    For positive integers \(m\) and \(n\), suppose that \(m^{\alpha} = n\). Show that \(\alpha = \frac{p}{q}\) with \(\gcd(p,q) = 1\) if and only if

    \[\begin{array}{ccccc} {m = \prod_{i=1}^{s} p_{i}^{k_{i}}}&{and}&{n = \prod_{i=1}^{s} p_{i}^{l_{i}}}&{with}&{\forall i pk_{i} = ql_{i}} \end{array} \nonumber\]

    Exercise \(\PageIndex{10}\)

    We develop the proof of Theorem 2.16 as it was given by Euler. We start by assuming that there is a finite list \(L\) of \(k\) primes. We will show in the following steps how that assumption leads to a contradiction. We order the list according to ascending order of magnitude of the primes. So \(L = \{p_{1}, p_{2}, \cdots, p_{k}\}\) where \(p_{1} = 2, p_{2} = 3, p_{3} = 5\), and so forth, upto the last prime \(p_{k}\).

    1. Show that \(\prod_{i=1}^{k} \frac{p_{i}}{p_{i}-1}\) is finite, say \(M\).
    2. Show that for \(r > 0\), \[\prod_{i=1}^{k} \frac{p_{i}}{p_{i}-1} = \prod_{i=1}^{k} \frac{1}{1-p_{i}^{-1}} > \prod_{i=1}^{k} \frac{1-p_{i}^{-r-1}}{1-p_{i}^{-1}} = \prod_{i=1}^{k} (\sum_{ j=0}^{r} p_{i}^{-j}) \nonumber\]
    3. Use the fundamental theorem of arithmetic to show that there is an \(\alpha (r) > 0\) such that \[\prod_{i=1}^{k} (\sum_{ j=0}^{r} p_{i}^{-j}) = \sum_{l=1}^{\alpha (r)} \frac{1}{l}+R \nonumber\], where \(R\) is a non-negative remainder.
    4. Show that for all \(K\) there is an \(r\) such that \(\alpha (r) > K\).
    5. Thus for any \(K\), there is an \(r\) such that \[\prod_{i=1}^{k} (\sum_{ j=0}^{r} p_{i}^{-j}) \ge \sum_{l=1}^{K} \frac{1}{l} \nonumber\]
    6. Conclude with a contradiction between a) and e). (Hint: the harmonic series \(\sum \frac{1}{l}\) diverges or see exercise 2.11 c).)

    Exercise \(\PageIndex{11}\)

    In this exercise we consider the Riemann zeta function for real values of \(s\) greater than \(1\).

    1. Show that for all \(x > -1\), we have \(\mbox{ln} (1+x) \le x\).
    2. Use Proposition 2.20 and a) to show that \[\mbox{ln} \zeta (s) \le \sum_{\mbox{p prime}} \frac{p^{-s}}{1-p^{-s}} \le \sum_{\mbox{p prime}} \frac{p^{-s}}{1-2^{-s}} \nonumber\]
    3. Use the following argument to show that \(\lim_{s \rightarrow 1} \zeta (s) = \infty\). \[\sum_{n=1}^{\infty} n^{-1} \ge \sum_{n=1}^{\infty} n^{-s} \ge \int_{1}^{\infty} x^{-s} dx \nonumber\] (For the last inequality, see Figure 4.)
    4. Show that b) and c) imply that \(\sum_{\mbox{p prime}} p^{-s}\) diverges as \(s \rightarrow 1\).
    5. Show that – in some sense – primes are more frequent than squares in the natural numbers. (Hint: \(\sum_{n=1}^{\infty} n^{-2}\) converges.)

    Screen Shot 2021-04-08 at 10.43.50 PM.png

    Figure 4. Proof that \(\sum_{n=1}^{\infty} f(n)\) is greater than \(\int_{1}^{\infty} x^{-s} dx\) if \(f\) is positive and decreasing.

    Exercise \(\PageIndex{12}\)

    Let \(E\) be the set of even numbers. Let \(a, c\) in \(E\), then \(c\) is divisible by \(a\) if there is \(ab \in E\) so that \(ab = c\). Define a prime \(p\) in \(E\) as a number in \(E\) such that there are no \(a\) and \(b\) in \(E\) with \(ab = p\).

    1. List the first \(30\) primes in \(E\).
    2. Does Euclid’s lemma hold in \(E\)? Explain.
    3. Factor 60 into primes (in \(E\)) in two different ways.

    Exercise \(\PageIndex{13}\)

    See exercise 2.12. Show that any number in \(E\) is a product of primes in \(E\). (Hint: follow the proof of Theorem 2.14, part (1).)

    Exercise \(\PageIndex{14}\)

    See exercise 2.12 which shows that unique factorization does not hold in \(E = \{2, 4, 6, \dots\}\). The proof of unique factorization uses Euclid’s lemma. In turn, Euclid’s lemma was a corollary of Be ́zout’s lemma, which depends on the division algorithm. Where exactly does the chain break down in this case?

    Exercise \(\PageIndex{15}\)

    Let \(L = \{p_{1}, p_{2}, \cdots\}\) be the list of all primes, ordered according ascending magnitude. Show that \(p_{n+1} \le \prod_{i=1}^{n} p_{i}\). (Hint: consider \(d\) as in the proof of Theorem 2.16, let \(p_{n+1}\) be the smallest prime divisor of \(d-1\).)

    A much stronger version of exercise 2.15 is the so-called Bertrand’s Postulate. That theorem says that for every \(n \ge 1\), there is a prime in \(\{n+1, \dots, 2n\}\). It was proved by Chebyshev. Subsequently the proof was simplified by Ramanujan and Erdo ̈s [18].

    Exercise \(\PageIndex{16}\)

    Let \(p\) and \(q\) primes greater than or equal to \(5\).

    1. Show that \(p = 12q+r\) where \(r \in \{1,5,7,11\}\). (The same holds for \(q\).)
    2. Show that \(24|p^2-q^2\). (Hint: use (a) to reduce this to \({4 \choose 2} = 6\) cases.)

    Exercise \(\PageIndex{17}\)

    A square full number is a number \(n > 1\) such that each prime factor occurs with a power at least \(2\). A square free number is a number \(n > 1\) such that each prime factor occurs with a power at most \(1\).

    1. If \(n\) is square full, show that there are positive integers \(a) and \(b\) such that \(n = a^{2}b^{3}\).
    2. Show that every integer greater than one is the product of a square free number and a square number.

    Exercise \(\PageIndex{18}\)

    Let \(L = \{p_{1}, p_{2},\dots\}\) be the list of all primes, ordered ac- cording ascending magnitude. The numbers \(E_{n} = 1 + \prod_{i=1}^{n} p_{i}\) are called Euclid numbers.

    1. Check the primality of \(E_{1}\) through \(E_{6}\).
    2. Show that \(E_{n} = _{4} 3\_. (Hint: \(E_{n}-1\) is twice an odd number.)
    3. Show that for \(n \le 3\) the decimal representation of \(E_{n}\) ends in a \(1\). (Hint: look at the factors of \(E_{n}\).)

    Exercise \(\PageIndex{19}\)

    Twin primes are a pair of primes of the form \(p\) and \(p+2\).

    1. Show that the product of two twin primes plus one is a square.
    2. Show that \(p > 3\), the sum of twin primes is divisible by \(12\). (Hint: see exercise 2.16)

    Exercise \(\PageIndex{20}\)

    Show that there arbitrarily large gaps between successive primes. More precisely, show that every integer in \(\{n!+2, n!+3, \dots n!+n\}\) is composite.

    The usual statement for the fundamental theorem of arithmetic includes only natural numbers \(n \in \mathbb{N}\) (i.e. not \(\mathbb{Z}\)) and the common proof uses induction on n. We review that proof in the next two problems.

    Exercise \(\PageIndex{21}\)

    1. Prove that \(2\) can be written as a product of primes.
    2. Let \(k > 2\). Suppose all numbers in \(\{1, 2, \dots k\}\) can be written as a product of primes (or \(1\)). Show that \(k+1\) is either prime or composite.
    3. If in (b), \(k+1\) is prime, then all numbers in \(\{1, 2, \dots k+1\}\) can be written as a product of primes (or \(1\)).
    4. If in (b), \(k+1\) is composite, then there is a divisor \(d \in \{2,\dots k\}\) such that \(k+1 = dd'\).
    5. Show that the hypothesis in (b) implies also in this case, all numbers in \(\{1, 2, \dots k+1\}\) can be written as a product of primes (or \(1\)).
    6. Use the above to formulate the inductive proof that all elements of \(\mathbb{N}\) can be written as a product of primes.

    Exercise \(\PageIndex{22}\)

    The set-up of the proof is the same as in exercise 2.21. Use induction on \(n\). We assume the result of that exercise.

    1. Show that \(n = 2\) has a unique factorization.
    2. Suppose that if for \(k > 2, \{2, \dots k\}\) can be uniquely factored. Then there are primes \(p_{i}\) and \(q_{i}\), not necessarily distinct, such that \[k+1 = \prod_{i=1}^{s} p_{i} = \prod_{i=1}^{r} q_{i} \nonumber\]
    3. Show that then \(p_{1}\) divides \(\prod_{i=1}^{r} q_{i}\) and so, Corollary 2.12 implies that there is a \(j \le r\) such that \(p_{1} = q_{j}\).
    4. Relabel the \(q_{i}'\)s, so that \(p_{1} = q_{1}\) and divide \(n\) by \(p_{1} = q_{1}\). Show that \[\frac{k+1}{q_{1}} = \prod_{i=2}^{s} p_{i} = \prod_{i=2}^{r} q_{i} \nonumber\]
    5. Show that the hypothesis in (b) implies that the remaining \(p_{i}\) equal the remaining \(q_{i}\). (Hint: \(\frac{k}{q_{1}} \le k\).)
    6. Use the above to formulate the inductive proof that all elements of \(\mathbb{N}\) can be uniquely factored as a product of primes.

    Here is a different characterization of gcd and lcm. We prove it as a corollary of the prime factorization theorem.

    Corollary 2.23

    1. A common divisor \(d > 0\) of \(a\) and \(b\) equals \(\gcd(a, b)\) if and only if every common divisor of \(a\) and \(b\) is a divisor of \(d\).
    2. Also, a common multiple \(d > 0\) of \(a\) and \(b\) equals \(\mbox{lcm} (a, b)\) if and only if every common multiple of \(a\) and \(b\) is a multiple of \(d\).

    Exercise \(\PageIndex{23}\)

    Use the characterization of \(\gcd(a, b)\) and \(\mbox{lcm} (a,b)\) given in the proof of Corollary 2.15 to prove Corollary 2.23.

    Exercise \(\PageIndex{24}\)

    1. Let \(p\) be a fixed prime. Show that the probability that two independently chosen integers in \(\{1,2,\cdots,n\}\) are divisible by \(p\) tends to \(\frac{1}{p^2}\) as \(n \rightarrow \infty\). Equivalently, the probability that they are not divisible by \(p\) tends to \(1-\frac{1}{p^2}\).
    2. Make the necessary assumptions, and show that the probability that two independently chosen integers in \(\{1,2,\cdots,n\}\) are not divisible by any prime tends to \(\prod_{p prime} (1-p^{-2})\). (Hint: you need to assume that the probabilities in 1. are independent and so they can be multiplied.)
    3. Show that from 2. and Euler's product formula, it follows that for 2 random (positive) integers \(a\) and \(b\) to have \(\gcd (a,b) = 1\) has probability \(1/\zeta (2) \approx 0.61\)
    4. Show that \(d >1\) integers \(\{a_{1}, a_{2}, \cdots, a_{d}\}\) that probability equals \(1/\zeta(d)\). (Hint: the reasoning is the same as in 1., 2., 3.)
    5. Show that for \(d \ge 2\): \[1 < \zeta (d) < 1+\int_{1}^{\infty} x^{-d} dx \nonumber\] For the last inequality, see Figure 5.
    6. Show that for large \(d\), the probability that \(\gcd (a_{1},a_{2},\cdots,a_{d}) = 1\) tends to \(1\).

    Screen Shot 2021-04-16 at 4.31.15 PM.png

    Figure 5. Proof that \(\sum_{n=1}^{\infty} f(n)\) (shaded) minus \(f(1)\) (shaded in blue) is less than \(\int_{1}^{\infty} x^{-s} dx\) is positive and decreasing.

    Exercise \(\PageIndex{25}\)

    This exercise in based on exercise 2.24.

    1. In the \(\{-4,\cdots, 4\}^2 \backslash (0,0)\) grid in \(\mathbb{Z}^{2}\), find out which propotion of the lattice points is visible from the origin, see Figure 6.
    2. Show that in a large grid, this propotion tends to \(1/\zeta (2)\).
    3. Show that as the dimension increases to infinity, the propotion of the lattice points \(\mathbb{Z}^{d}\) that are visible from the origin, increases to 1.

    Screen Shot 2021-04-16 at 4.36.53 PM.png

    Figure 6. The origin is marked by “\(\times\)”. The red dots are visible from \(\times\); between any blue dot and \(\times\) there is a red dot. The picture shows exactly one quarter of \(\{-4,\cdots ,4\}^{2} \backslash (0,0) \subset \mathbb{Z}^2\).


    This page titled 2.6: Exercises is shared under a CC BY-NC license and was authored, remixed, and/or curated by J. J. P. Veerman (PDXOpen: Open Educational Resources) .

    • Was this article helpful?