跳到主要内容

1 Real Numbers and Limits of Sequences

1.1 Number Systems

Natural number

The Peano axioms describe the arithmetic properties of the set of natural numbers N\mathbb{N}. The formal definition is as follows:

  1. There exists an element 0N0 \in \mathbb{N}.

  2. There exists a function

    S:NN\begin{equation*} S : \mathbb{N} \to \mathbb{N} \end{equation*}

    called the successor function, such that for every nNn \in \mathbb{N}, S(n)S(n) represents the next natural number after nn.

  3. Zero is not the successor of any number

    nN,  S(n)0\begin{equation*} \forall n \in \mathbb{N},\; S(n) \neq 0 \end{equation*}
  4. Successor function is injective (one-to-one)

    m,nN,  S(m)=S(n)    m=n\begin{equation*} \forall m, n \in \mathbb{N},\; S(m) = S(n) \implies m = n \end{equation*}
  5. Principle of mathematical induction. For any property P(n)P(n) defined on N\mathbb{N}: If

    P(0) is true, and nN,  P(n)P(S(n)),\begin{equation*} P(0) \text{ is true, and } \forall n \in \mathbb{N},\; P(n) \Rightarrow P(S(n)), \end{equation*}

    then

    nN,  P(n) is true.\begin{equation*} \forall n \in \mathbb{N},\; P(n) \text{ is true.} \end{equation*}

Addition

The addition +:N×NN+ : \mathbb{N} \times \mathbb{N} \to \mathbb{N} is defined recursively as follows:

a+0=a\begin{equation*} a + 0 = a \end{equation*} a+S(b)=S(a+b)\begin{equation*} a + S(b) = S(a + b) \end{equation*}

Multiplication

The multiplication ×:N×NN\times : \mathbb{N} \times \mathbb{N} \to \mathbb{N} is defined recursively as follows:

a×0=0\begin{equation*} a \times 0 = 0 \end{equation*} a×S(b)=a×b+a\begin{equation*} a \times S(b) = a \times b + a \end{equation*}

Order Relation

For all m,nNm, n \in \mathbb{N}:

m<S(n)    (m<n) or (m=n).\begin{equation*} m < S(n) \iff (m < n) \text{ or } (m = n). \end{equation*}

This defines << recursively based on the successor structure.

Theorems

  1. a+(b+c)=(a+b)+ca + (b + c) = (a + b) + c.

  2. a+0=0+a=aa + 0 = 0 + a = a.

  3. a+b=b+aa + b = b + a.

  4. (a+b)×c=a×c+b×c(a + b)\times c = a \times c + b \times c.

  5. a×1=1×a=aa \times 1 = 1 \times a = a.

  6. a×b=b×aa \times b = b \times a.

  7. a×(b×c)=(a×b)×ca \times (b \times c) = (a \times b) \times c.

Integer

Integers can be defined from natural numbers as equivalence classes of ordered pairs of natural numbers:

Z={(a,b)a,bN}\begin{equation*} \mathbb{Z} = \{ (a, b) \mid a, b \in \mathbb{N} \} \end{equation*}

We define an equivalence relation:

(a,b)(c,d)    a+d=b+c\begin{equation*} (a, b) \sim (c, d) \iff a + d = b + c \end{equation*}
  • Each equivalence class [(a,b)][(a, b)] represents the difference between aa and bb. Thus we have: 0=[(0,0)],1=[(1,0)],1=[(0,1)],ab=[(a,b)].\begin{equation*} \begin{aligned} 0 &= [(0, 0)], \\ 1 &= [(1, 0)], \\ -1 &= [(0, 1)], \\ a - b &= [(a, b)]. \end{aligned} \end{equation*}

Addition

[(a,b)]+[(c,d)]=[(a+c,b+d)]\begin{equation*} [(a, b)] + [(c, d)] = [(a + c,\, b + d)] \end{equation*}

Multiplication

[(a,b)]×[(c,d)]=[(ac+bd,ad+bc)]\begin{equation*} [(a, b)] \times [(c, d)] = [(ac + bd,\, ad + bc)] \end{equation*}

These definitions are well-defined, meaning the result does not depend on the choice of representatives of the equivalence classes.

Order Relation

[(a,b)]<[(c,d)]    a+d<b+c.\begin{equation*} [(a,b)] < [(c,d)] \iff a + d < b + c. \end{equation*}

Theorems

  1. a+b=b+a,a×b=b×aa + b = b + a, \qquad a \times b = b \times a.

  2. a+(b+c)=(a+b)+c,a×(b×c)=(a×b)×ca + (b + c) = (a + b) + c, \qquad a \times (b \times c) = (a \times b) \times c.

  3. a+0=a,a×1=aa + 0 = a, \qquad a \times 1 = a.

  4. aZ,  (a)Z such that a+(a)=0\forall a \in \mathbb{Z}, \; \exists (-a) \in \mathbb{Z} \text{ such that } a + (-a) = 0.

  5. a×(b+c)=a×b+a×ca \times (b + c) = a \times b + a \times c.

Rational Number

The set of rational numbers Q\mathbb{Q} is constructed from the integers Z\mathbb{Z} as the set of equivalence classes of ordered pairs of integers:

Q={(a,b)aZ,  bZ{0}}\begin{equation*} \mathbb{Q} = \{ (a, b) \mid a \in \mathbb{Z},\; b \in \mathbb{Z} \setminus \{0\} \} \end{equation*}

Each pair (a,b)(a,b) intuitively represents the fraction ab\dfrac{a}{b}.

We define an equivalence relation:

(a,b)(c,d)    ad=bc\begin{equation*} (a,b) \sim (c,d) \iff ad = bc \end{equation*}

Each equivalence class [(a,b)][(a,b)] corresponds to a rational number. We write:

ab=[(a,b)],where b0\begin{equation*} \dfrac{a}{b} = [(a,b)], \quad \text{where } b \neq 0 \end{equation*}

Addition

ab+cd=ad+bcbd\begin{equation*} \dfrac{a}{b} + \dfrac{c}{d} = \dfrac{ad + bc}{bd} \end{equation*}

Multiplication

ab×cd=acbd\begin{equation*} \dfrac{a}{b} \times \dfrac{c}{d} = \dfrac{ac}{bd} \end{equation*}

Additive Inverse

ab=ab\begin{equation*} -\dfrac{a}{b} = \dfrac{-a}{b} \end{equation*}

Multiplicative Inverse

(ab)1=ba,a0\begin{equation*} \left(\dfrac{a}{b}\right)^{-1} = \dfrac{b}{a}, \quad a \neq 0 \end{equation*}

Order relation

To compare two rational numbers, let

ab,  cdQ,b>0,  d>0.\begin{equation*} \dfrac{a}{b}, \; \dfrac{c}{d} \in \mathbb{Q}, \quad b > 0, \; d > 0. \end{equation*}

We define the order relation << on Q\mathbb{Q} by:

ab<cd    ad<bc.\begin{equation*} \dfrac{a}{b} < \dfrac{c}{d} \iff ad < bc. \end{equation*}

Theorems

  1. a+b=b+a,a×b=b×aa+b = b+a, \qquad a\times b = b\times a.

  2. a+(b+c)=(a+b)+c,a×(b×c)=(a×b)×ca+(b+c) = (a+b)+c, \qquad a\times(b\times c) = (a\times b)\times c.

  3. a+0=a,a×1=aa+0 = a, \qquad a\times1 = a.

  4. aQ,  (a) such that a+(a)=0\forall a \in \mathbb{Q},\; \exists (-a) \text{ such that } a+(-a)=0.

  5. aQ{0},  a1 such that a×a1=1\forall a \in \mathbb{Q}\setminus\{0\},\; \exists a^{-1} \text{ such that } a\times a^{-1}=1.

  6. a×(b+c)=a×b+a×ca\times(b+c) = a\times b + a\times c.

Real number

A Dedekind cut in Q\mathbb{Q} is a partition (A,B)(A, B) of Q\mathbb{Q} satisfying:

(1)A,BQ,  AB=Q,  AB=,(2)aA,  bB,  a<b,(3)A has no greatest element.\begin{equation*} \begin{aligned} (1)\quad & A, B \subset \mathbb{Q}, \; A \cup B = \mathbb{Q}, \; A \cap B = \varnothing, \\[4pt] (2)\quad & \forall a \in A,\; \forall b \in B,\; a < b, \\[4pt] (3)\quad & A \text{ has no greatest element.} \end{aligned} \end{equation*}

Each real number is identified with one such cut AA, which intuitively represents "all rationals less than that real number."

The set of all Dedekind cuts is denoted by:

R={AQA is a Dedekind cut}.\begin{equation*} \mathbb{R} = \{ A \subset \mathbb{Q} \mid A \text{ is a Dedekind cut} \}. \end{equation*}
  • For example:
2={qQq<0 or q2<2}.\begin{equation*} \sqrt{2} = \{ q \in \mathbb{Q} \mid q < 0 \text{ or } q^2 < 2 \}. \end{equation*}

Addition

A+B={a+baA,bB}.\begin{equation*} A + B = \{ a + b \mid a \in A,\, b \in B \}. \end{equation*}

Multiplication

(for positive cuts)

A×B={a×baA,bB,  a,b>0}.\begin{equation*} A \times B = \{ a \times b \mid a \in A,\, b \in B,\; a,b > 0 \}. \end{equation*}

Negatives and general cases are defined symmetrically by extending signs.

Order Relation

Define:

A<B    AB.\begin{equation*} A < B \iff A \subsetneq B. \end{equation*}

This relation extends the usual order on Q\mathbb{Q} and makes (R,<)(\mathbb{R}, <) a totally ordered set.

Theorems

  1. a+b=b+a,a×b=b×aa+b = b+a, \qquad a\times b = b\times a.

  2. a+(b+c)=(a+b)+c,a×(b×c)=(a×b)×ca+(b+c) = (a+b)+c, \qquad a\times(b\times c) = (a\times b)\times c.

  3. a+0=a,a×1=aa+0 = a, \qquad a\times1 = a.

  4. aQ,  (a) such that a+(a)=0\forall a \in \mathbb{Q},\; \exists (-a) \text{ such that } a+(-a)=0.

  5. aQ{0},  a1 such that a×a1=1\forall a \in \mathbb{Q}\setminus\{0\},\; \exists a^{-1} \text{ such that } a\times a^{-1}=1.

  6. a×(b+c)=a×b+a×ca\times(b+c) = a\times b + a\times c.

  7. The set of rational numbers Q\mathbb{Q} is dense in R\mathbb{R}:

a,bR,  a<brQ such that a<r<b.\begin{equation*} \forall a,b \in \mathbb{R},\; a < b \Rightarrow \exists r \in \mathbb{Q} \text{ such that } a < r < b. \end{equation*}

1.2 Sequences and Convergent Sequences

Convergent Sequence

Let {an}\{a_n\} be a sequence and aa be a real number. If for any given ε>0\varepsilon > 0, there exists an NNN \in \mathbb{N}^* such that whenever n>Nn > N, we have:

ana<ε\begin{equation*} |a_n - a| < \varepsilon \end{equation*}

then the sequence {an}\{a_n\} is said to have the limit aa as nn tends to infinity, denoted as:

limnan=a\begin{equation*} \lim_{n \to \infty} a_n = a \end{equation*}

It can also be briefly denoted as anaa_n \to a (nn \to \infty). We also say that the sequence {an}\{a_n\} converges to aa. A sequence that has a limit is called a convergent sequence; a sequence that does not converge is called a divergent sequence.

Examples

  1. α>0\alpha > 0, limn1nα=0. \lim\limits_{n \to \infty} \frac{1}{n^\alpha} = 0.
  2. limnn1/n=1. \lim\limits_{n \to \infty} n^{1/n} = 1.

Theorem 1.1. If the sequence {an}\{a_n\} converges, then it has only one limit. That is to say, the limit of a convergent sequence is unique.

Bounded Sequence

Let {an}\{a_n\} be a sequence. If there exists a real number AA, such that anAa_n \leqslant A holds for all nNn \in \mathbb{N}^*, then {an}\{a_n\} is said to be bounded above, and AA is an upper bound of this sequence.

Similarly, we can define a sequence that is bounded below.

If the sequence {an}\{a_n\} has both a lower bound and an upper bound, then it is called a bounded sequence.

Theorem 1.2. Convergent sequences are bounded.

Subsequence

Let {an}\{a_n\} be a sequence, and let kiNk_i \in \mathbb{N}^* (i=1,2,3,i=1, 2, 3, \dots) satisfy k1<k2<k3<k_1 < k_2 < k_3 < \dots. Then the sequence {akn}\{a_{k_n}\} is called a subsequence of {an}\{a_n\}.

Theorem 1.3. Let the limit of a convergent sequence {an}\{a_n\} be aa. Then any subsequence of {an}\{a_n\} also converges to aa.

Corollary 1. A necessary and sufficient condition for the sequence {an}\{a_n\} to converge is that both its even subsequence {a2n}\{a_{2n}\} and its odd subsequence {a2n1}\{a_{2n-1}\} converge, and have the same limit.

Theorem 1.4. Let {an}\{a_n\} and {bn}\{b_n\} both be convergent sequences. Then {an+bn}\{a_n + b_n\} and {anbn}\{a_n b_n\} are also convergent sequences. If limnbn0\lim\limits_{n \to \infty} b_n \neq 0, then {an/bn}\{a_n / b_n\} is also convergent, and:

(1)limn(an±bn)=limnan±limnbn.\text{(1)}\quad \lim\limits_{n \to \infty} (a_n \pm b_n) = \lim\limits_{n \to \infty} a_n \pm \lim\limits_{n \to \infty} b_n.

(2)limnanbn=limnanlimnbn.\text{(2)}\quad \lim\limits_{n \to \infty} a_n b_n = \lim\limits_{n \to \infty} a_n \cdot \lim\limits_{n \to \infty} b_n.

(3)limnanbn=limnanlimnbn,where limnbn0.\text{(3)}\quad \lim\limits_{n \to \infty} \frac{a_n}{b_n} = \frac{\lim\limits_{n \to \infty} a_n}{\lim\limits_{n \to \infty} b_n}, \quad \text{where } \lim\limits_{n \to \infty} b_n \neq 0.

Infinitesimal Sequence

If the limit of a convergent sequence {an}\{a_n\} is equal to 0, then this sequence is called an infinitesimal sequence, or simply referred to as an infinitesimal.

Theorem 1.5. Let

anbncn(nN)\begin{equation*} a_n \leqslant b_n \leqslant c_n \quad (n \in \mathbb{N}^*) \end{equation*}

If limnan=limncn=a\lim\limits_{n \to \infty} a_n = \lim\limits_{n \to \infty} c_n = a, then

limnbn=a\begin{equation*} \lim_{n \to \infty} b_n = a \end{equation*}

Examples

  1. a>1,kNa > 1, k \in \mathbb{N}^*, limnnkan=0.\lim\limits_{n \to \infty}\frac{n^k}{a^n} = 0.
  2. limn1n2+1+1n2+2++1n2+n=1\lim\limits_{n \to \infty} \frac{1}{\sqrt{n^2+1}} + \frac{1}{\sqrt{n^2+2}} + \cdots + \frac{1}{\sqrt{n^2+n}} = 1.

Infinite Sequence

If the sequence {an}\{a_n\} satisfies the condition: for any positive number AA, there exists an NNN \in \mathbb{N}^* such that whenever n>Nn > N, we have an>Aa_n > A, then the sequence {an}\{a_n\} is said to tend to ++\infty (positive infinity), denoted as

limnan=+.\begin{equation*} \lim_{n \to \infty} a_n = +\infty. \end{equation*}

If for any positive number AA, there exists an NNN \in \mathbb{N}^* such that whenever n>Nn > N, we have an<Aa_n < -A, then the sequence {an}\{a_n\} is said to tend to -\infty (negative infinity), denoted as

limnan=.\begin{equation*} \lim_{n \to \infty} a_n = -\infty. \end{equation*}

If limnan=+\lim\limits_{n \to \infty} |a_n| = +\infty, then the sequence {an}\{a_n\} is said to tend to \infty, denoted as limnan=\lim\limits_{n \to \infty} a_n = \infty.

Regardless of which of the three cases

limnan=+,limnan=,limnan=\begin{equation*} \lim_{n \to \infty} a_n = +\infty, \quad \lim_{n \to \infty} a_n = -\infty, \quad \lim_{n \to \infty} a_n = \infty \end{equation*}

holds, the sequence {an}\{a_n\} is called infinity (or an infinite sequence).

1.3 Monotonic Sequences

Monotonic Sequence

If the sequence {an}\{a_n\} satisfies

anan+1(n=1,2,),\begin{equation*} a_n \leqslant a_{n+1} \quad (n = 1, 2, \dots), \end{equation*}

then this sequence is called an increasing sequence; if {an}\{a_n\} satisfies

anan+1(n=1,2,),\begin{equation*} a_n \geqslant a_{n+1} \quad (n = 1, 2, \dots), \end{equation*}

then this sequence is called a decreasing sequence. If the above two inequalities are strict, i.e., an<an+1a_n < a_{n+1} (or an>an+1a_n > a_{n+1}) (n=1,2,n=1, 2, \dots), then this sequence is called strictly increasing (or strictly decreasing).

Increasing or decreasing sequences are collectively called monotonic sequences.

Theorem 1.6. A monotonic and bounded sequence must have a limit.

Theorem 1.7. For nNn \in \mathbb{N}^*, let the sequence {an}\{a_n\} be defined by:

an=1+12α++1nα\begin{equation*} a_n = 1 + \frac{1}{2^\alpha} + \cdots + \frac{1}{n^\alpha} \end{equation*}

where α>1\alpha > 1. Then the sequence {an}\{a_n\} is convergent.

Theorem 1.8. Let In=[an,bn]I_n = [a_n, b_n] (nNn \in \mathbb{N}^*), and let

I1I2I3InIn+1.\begin{equation*} I_1 \supset I_2 \supset I_3 \supset \cdots \supset I_n \supset I_{n+1} \supset \cdots. \end{equation*}

If the length of this sequence of intervals In=bnan0|I_n| = b_n - a_n \to 0 (nn \to \infty), then the intersection

n=1In\begin{equation*} \bigcap_{n=1}^{\infty} I_n \end{equation*}

contains a unique point.

1.4 The Base of Natural Logarithm e

Definition of e

Let two sequences be defined as

an=(1+1n)n,bn=(1+1n)n+1.\begin{equation*} a_n = \left(1 + \frac{1}{n}\right)^n, \quad b_n = \left(1 + \frac{1}{n}\right)^{n+1}. \end{equation*}

Then:

  • ana_n is monotonically increasing.
  • bnb_n is monotonically decreasing.
  • For all nn, an<bna_n < b_n.
  • Therefore, both sequences converge and have the same limit.

We define this common limit as

e=limn+(1+1n)n.\begin{equation*} e = \lim_{n \to +\infty} \left(1 + \frac{1}{n}\right)^n. \end{equation*}

Theorem 1.9. Let the sequence {sn}\{s_n\} be defined as:

sn=1+11!+12!++1n!(nN).\begin{equation*} s_n = 1 + \frac{1}{1!} + \frac{1}{2!} + \cdots + \frac{1}{n!} \quad (n \in \mathbb{N}^*). \end{equation*}

Then for any positive integer nn, the error in approximating the base of the natural logarithm ee by sns_n satisfies the following inequality:

0<esn1n!n(nN).\begin{equation*} 0 < e - s_n \leqslant \frac{1}{n! n} \quad (n \in \mathbb{N}^*). \end{equation*}

1.5 Fundamental Sequences and Cauchy Convergence Criterion

Cauchy Sequence

Let {an}\{a_n\} be a real sequence. For any given ε>0\varepsilon > 0, if there exists an NNN \in \mathbb{N}^* such that whenever m,nNm, n \in \mathbb{N}^* and m,n>Nm, n > N, we have

aman<ε,\begin{equation*} | a_m - a_n | < \varepsilon, \end{equation*}

then the sequence {an}\{a_n\} is called a fundamental sequence or a Cauchy sequence.

Lemma. Every sequence contains a monotonic subsequence.

Theorem 1.10. From any bounded sequence, a convergent subsequence can always be selected.

Theorem 1.11. A necessary and sufficient condition for a sequence to converge is that it is a fundamental sequence.

1.6 Supremum and Infimum

Supremum

Let EE be a non-empty set bounded above. A real number β\beta satisfies the following two conditions:

(1)\text{(1)}\quadFor any xEx \in E, we have xβx \leqslant \beta;

(2)\text{(2)}\quadFor any given ε>0\varepsilon > 0, there must exist an xεEx_{\varepsilon} \in E such that xε>βεx_{\varepsilon} > \beta - \varepsilon.

In this case, β\beta is called the supremum of the set EE, denoted as β=supE\beta = \sup E.

Infimum

Let EE be a non-empty set bounded below. A real number α\alpha satisfies the following two conditions:

(1)\text{(1)}\quadFor any xEx \in E, we have xαx \geqslant \alpha;

(2)\text{(2)}\quadFor any given ε>0\varepsilon > 0, there must exist a yεEy_{\varepsilon} \in E such that yε<α+εy_{\varepsilon} < \alpha + \varepsilon.

In this case, α\alpha is called the infimum of the set EE, denoted as α=infE\alpha = \inf E.

Theorem 1.12. A non-empty set bounded above must have a supremum; a non-empty set bounded below must have an infimum.

Theorem 1.13. Let [a,b][a, b] be a finite closed interval, and let {Iλ}\{I_\lambda\} be an open cover of it. Then a finite number of members (open intervals) can necessarily be selected from this family of open intervals, such that the family formed by these finite open intervals is still an open cover of [a,b][a, b].

1.7 Limit Superior and Limit Inferior

Limit Superior and Limit Inferior

Let {an}\{a_n\} be a sequence, and let EE be the set composed of all limit points of {an}\{a_n\}. Let

a=supE,a=infE,\begin{equation*} a^* = \sup E, \quad a_* = \inf E, \end{equation*}

then aa^* and aa_* are called the limit superior and limit inferior of the sequence {an}\{a_n\}, respectively, denoted as

a=lim supnan,a=lim infnan.\begin{equation*} a^* = \limsup_{n \to \infty} a_n, \quad a_* = \liminf_{n \to \infty} a_n. \end{equation*}

Theorem 1.14. Let {an}\{a_n\} be a sequence. Then:

(1)aE\text{(1)}\quad a^* \in E;

(2)\text{(2)}\quadIf x>ax > a^*, then there exists an NNN \in \mathbb{N}^* such that whenever nNn \geqslant N, we have an<xa_n < x;

(3)a\text{(3)}\quad a^* is the unique number satisfying the first two properties.

Theorem 1.15. For the sequence {an}\{a_n\}, define αn=infknak\alpha_n = \inf\limits_{k \geqslant n} a_k and βn=supknak\beta_n = \sup\limits_{k \geqslant n} a_k. Then:

(1){αn}\text{(1)}\quad \{\alpha_n\} is an increasing sequence, and {βn}\{\beta_n\} is a decreasing sequence;

(2)limnαn=a,limnβn=a\text{(2)}\quad \lim\limits_{n \to \infty} \alpha_n = a_*, \quad \lim\limits_{n \to \infty} \beta_n = a^*.

Stolz Theorem

Theorem 1.16 (\frac{\infty}{\infty} Type). Let {bn}\{b_n\} be a strictly increasing sequence tending to ++\infty. If

limnanan1bnbn1=A,\begin{equation*} \lim_{n \to \infty} \frac{a_n - a_{n-1}}{b_n - b_{n-1}} = A, \end{equation*}

then

limnanbn=A,\begin{equation*} \lim_{n \to \infty} \frac{a_n}{b_n} = A, \end{equation*}

where AA can be ++\infty or -\infty.

Theorem 1.17 (00\frac{0}{0} Type). Let {an}\{a_n\} and {bn}\{b_n\} be two sequences converging to 0. If {bn}\{b_n\} is strictly decreasing and

limnanan1bnbn1=A,\begin{equation*} \lim_{n \to \infty} \frac{a_n - a_{n-1}}{b_n - b_{n-1}} = A, \end{equation*}

then

limnanbn=A,\begin{equation*} \lim_{n \to \infty} \frac{a_n}{b_n} = A, \end{equation*}

where AA can be a finite number, ++\infty, or -\infty.