The Real and Complex Number Systems
The Real and Complex Number Systems
Section titled “The Real and Complex Number Systems”Integers
Section titled “Integers”Problem 1.1: No Largest Prime
Section titled “Problem 1.1: No Largest Prime”Problem. Prove that there is no largest prime. (A proof was known to Euclid.)
Strategy: Use proof by contradiction. Assume there exists a largest prime , then consider . Since is either prime or has a prime factor greater than , this contradicts the assumption.
Solution: We will prove this by contradiction. Assume there exists a largest prime number, call it .
Consider the number , where is the factorial of .
Since is divisible by all integers from to , the number is not divisible by any prime number less than or equal to .
Now, is either prime or composite:
- If is prime, then , contradicting our assumption that is the largest prime.
- If is composite, then has a prime factor . Since is not divisible by any prime , we must have . This again contradicts our assumption that is the largest prime.
In both cases, we reach a contradiction. Therefore, our assumption that there exists a largest prime is false, and there must be infinitely many prime numbers.
∎
Problem 1.2: Algebraic Identity
Section titled “Problem 1.2: Algebraic Identity”Problem. If is a positive integer, prove the algebraic identity:
Strategy: Expand the right-hand side and show it equals the left-hand side by distributing and observing that most terms cancel out.
Solution: We can prove this identity by expanding the right-hand side and showing it equals the left-hand side.
Let’s expand the sum:
Now distribute :
Notice that most terms cancel out:
Alternatively, we can use the geometric series formula. Let . Then:
Therefore, .
∎
Problem 1.3: Mersenne Primes
Section titled “Problem 1.3: Mersenne Primes”Problem. If is prime, prove that is prime. A prime of the form , where is prime, is called a Mersenne prime.
Strategy: Prove the contrapositive: if is composite, then is composite. Use the identity from Problem 1.2 to factor when with .
Solution: We will prove the contrapositive: if is composite, then is composite.
Let where are integers. Then:
Using the identity from Problem 1.2 with and :
Since , we have . Also, since , the second factor is greater than 1. Therefore, is the product of two integers greater than 1, making it composite.
This proves that if is prime, then must be prime.
∎
Problem 1.4: Fermat Primes
Section titled “Problem 1.4: Fermat Primes”Problem. If is prime, prove that is a power of . A prime of the form is called a Fermat prime. Hint: Use Exercise 1.2.
Strategy: Prove the contrapositive: if is not a power of , then is composite. When has an odd factor, use the identity from Problem 1.2 to factor .
Solution: We will prove the contrapositive: if is not a power of , then is composite.
If is not a power of , then has an odd factor greater than . Let where is odd and .
Then:
Since is odd, we can use the identity from Problem 1.2 with and :
Since is odd, , so:
Since , both factors are greater than , making composite.
Therefore, if is prime, then must be a power of .
∎
Problem 1.5: Fibonacci Numbers Formula
Section titled “Problem 1.5: Fibonacci Numbers Formula”Problem. The Fibonacci numbers are defined by the recursion formula , with . Prove that , where and are the roots of the equation .
Strategy: Use strong mathematical induction. Verify base cases for and , then use the inductive hypothesis and the key property that and to establish the inductive step.
Solution: Let the proposition be . The roots of are and . A key property of these roots is that they satisfy and .
Base cases: For :
For :
The base cases hold.
Inductive step: Assume is true for all integers , where . We will show is true. By the definition of the Fibonacci sequence, . Using the inductive hypothesis for and :
Using the property that and :
This is . By the principle of strong induction, the formula is true for all positive integers .
∎
Problem 1.6: Well-Ordering Principle
Section titled “Problem 1.6: Well-Ordering Principle”Problem. Prove that every nonempty set of positive integers contains a smallest member. This is called the well-ordering principle.
Strategy: Use proof by contradiction combined with mathematical induction. Assume there exists a nonempty set with no smallest member, then use induction to show this leads to the set being empty.
Solution: We will prove this by contradiction. Let be a nonempty set of positive integers that has no smallest member. Let be the proposition that the integer is not in . We will use induction to show that is true for all positive integers .
Base case: For : If , then would be the smallest member of (since contains only positive integers). But we assumed has no smallest member. So cannot be in . Thus, is true.
Inductive step: Assume that is true for all positive integers . This means that none of the integers are in . Now consider the integer . If , then from our inductive hypothesis (that are not in ), would be the smallest member of . This contradicts our initial assumption that has no smallest member. Therefore, cannot be in . Thus, is true.
By the principle of strong induction, is true for all positive integers . This means that no positive integer is in , which implies that is an empty set. This contradicts our initial assumption that is a nonempty set. Therefore, the original assumption must be false, and every nonempty set of positive integers must contain a smallest member.
∎
Rational and Irrational Numbers
Section titled “Rational and Irrational Numbers”Definitions and Theorems
Section titled “Definitions and Theorems”Theorem (Geometric Series Formula).
For , the infinite geometric series converges:
For a finite geometric series:
Importance: The geometric series is one of the most fundamental infinite series in mathematics. It provides the foundation for understanding power series, Taylor series, and many other important series expansions. This formula is essential for calculus, analysis, and many applications in physics and engineering.
Theorem (Terminating Decimal Criterion).
A rational number has a terminating decimal expansion if and only if the prime factorization of contains only powers of and .
Importance: This theorem provides a precise criterion for determining when a rational number has a finite decimal representation. It connects number theory (prime factorization) to decimal arithmetic, making it easy to predict the behavior of decimal expansions. This result is essential for understanding decimal representations and their applications.
Theorem (Irrationality of Square Roots).
If is a positive integer that is not a perfect square, then is irrational.
Importance: This theorem provides a large class of irrational numbers and is fundamental for understanding the relationship between algebra and geometry. It shows that many naturally occurring numbers (like , , etc.) are irrational, which is essential for geometry, algebra, and many areas of mathematics. The proof technique used here is a classic example of proof by contradiction.
Theorem (Density of Rationals and Irrationals).
Between any two real numbers, there exist both rational and irrational numbers.
Importance: This theorem shows that both rational and irrational numbers are densely distributed throughout the real number line. This property is fundamental for understanding the structure of the real numbers and is essential for many results in analysis, topology, and approximation theory. It demonstrates that the real number system is rich and complex.
Problem 1.7: Decimal Expansion to Rational
Section titled “Problem 1.7: Decimal Expansion to Rational”Problem. Find the rational number whose decimal expansion is .
Strategy: Use an algebraic method by multiplying by powers of 10 to shift the decimal point and eliminate the repeating part, then solve for the unknown fraction.
Solution: We can use an algebraic method to find the equivalent fraction. Let be the rational number. The goal is to manipulate the equation to eliminate the repeating decimal part. The repeating digit ‘4’ begins at the third decimal place.
First, multiply by to move the non-repeating part to the left of the decimal point: Next, multiply by to shift the decimal point past the first repeating digit: Now, subtract the first equation from the second. This will cancel the infinite repeating tail.
Finally, solve for : Therefore, the rational number is **\frac{301**{900}}.
∎
Problem 1.8: Decimal Expansion Ending in Zeroes
Section titled “Problem 1.8: Decimal Expansion Ending in Zeroes”Problem. Prove that the decimal expansion of will end in zeroes (or in nines) if and only if is a rational number whose denominator is of the form , where and are nonnegative integers.
Strategy: Prove both directions of the if-and-only-if statement. Show that rational numbers with denominators of the form have terminating decimal expansions, and conversely that terminating decimals correspond to such rational numbers.
Solution: We need to prove both directions of this statement.
Forward direction: If is rational with denominator of the form , then its decimal expansion terminates.
Let where for some nonnegative integers .
We can write
This shows that can be written as a fraction with denominator a power of 10, which means its decimal expansion terminates.
Reverse direction: If the decimal expansion of terminates, then is rational with denominator of the form .
Let have a terminating decimal expansion. Then can be written as for some integer and nonnegative integer .
Since , we have , which is of the required form.
Note about ending in nines: If a decimal expansion ends in nines (e.g., ), this is equivalent to the next terminating decimal. For example, . This is because .
Therefore, both terminating decimals and those ending in nines correspond to rational numbers with denominators of the form .
∎
Problem 1.9: Irrationality of
Section titled “Problem 1.9: Irrationality of 2+3\sqrt{2} + \sqrt{3}2+3”Problem. Prove that is irrational.
Strategy: Use proof by contradiction. Assume is rational, then square both sides to eliminate the square roots. This leads to being rational, which is false.
Solution: We will prove this by contradiction. Assume that is rational, say where are integers with no common factors.
Then:
This shows that is rational, which is a contradiction since is irrational.
To see why is irrational, suppose where are integers with no common factors. Then:
This means is divisible by 6, so must be divisible by 6. Let . Then:
This means is divisible by 6, so must also be divisible by 6. But this contradicts our assumption that and have no common factors.
Therefore, is irrational, and consequently is irrational.
∎
Problem 1.10: Rational Functions of Irrational Numbers
Section titled “Problem 1.10: Rational Functions of Irrational Numbers”Problem. If are rational and if is irrational, prove that is usually irrational. When do exceptions occur?
Strategy: Assume the expression is rational and solve for the conditions under which this can happen. Use the fact that if a rational expression equals a rational number, then the coefficients must satisfy certain relationships.
Solution: We need to analyze when is rational, given that is irrational and are rational.
Let’s assume that where are integers with no common factors.
Then:
Since is irrational and the right-hand side is rational, we must have and .
This gives us:
From the first equation: From the second equation:
Therefore, we have:
This means , or equivalently, .
Conclusion: The expression is rational if and only if .
Exceptions occur when: , which means the numerator and denominator are proportional, making the fraction rational regardless of the value of .
Examples:
- If , then , so for all .
- If , then , so for all .
∎
Problem 1.11: Irrational Numbers Between 0 and x
Section titled “Problem 1.11: Irrational Numbers Between 0 and x”Problem. Given any real , prove that there is an irrational number between and .
Strategy: Construct an irrational number between and by considering two cases: when is irrational (use ) and when is rational (use ).
Solution: We will construct an irrational number between and for any positive real number .
Case 1: If is irrational, then is irrational and lies between and .
To see why is irrational, suppose it were rational. Then for some integers , which would mean , making rational, a contradiction.
Case 2: If is rational, let where are positive integers.
Consider the number .
Since is irrational, is irrational (if were rational, then would be rational, a contradiction).
Also, since , we have .
Therefore, is an irrational number between and .
Alternative construction: For any positive real , we can also use . Since is irrational and greater than , we have , and is irrational.
∎
Problem 1.12: Fraction Between Two Fractions
Section titled “Problem 1.12: Fraction Between Two Fractions”Problem. If with , prove that lies between and .
Strategy: Prove both inequalities and by cross-multiplying and using the given condition .
Solution: We need to prove that .
Let’s prove both inequalities:
First inequality:
Cross-multiplying:
Since , we have , so this inequality holds.
Second inequality:
Cross-multiplying:
Again, since , we have , so this inequality also holds.
Therefore, lies between and .
Geometric interpretation: This result is known as the “mediant” of two fractions. If we think of fractions as points on a line, the mediant lies between the two original fractions and .
∎
Problem 1.13: Between Fractions
Section titled “Problem 1.13: 2\sqrt{2}2 Between Fractions”Problem. Let and be positive integers. Prove that always lies between the two fractions and . Which fraction is closer to ?
Strategy: Analyze the relationship between the two fractions and by examining their difference. Consider two cases based on whether is less than or greater than , then compare the distances to determine which fraction is closer.
Solution: Let’s first establish the ordering of the two fractions by examining their difference:
The sign of this difference depends on the sign of , which relates to .
**Case 1: \frac{a**{b} < \sqrt{2}}. This means , so , and . Thus, . We need to show that .
Since , this simplifies to , which is true by our case assumption. Thus, .
**Case 2: \frac{a**{b} > \sqrt{2}}. This means , and . Thus, . A similar calculation shows that . Therefore, . In both cases, lies between the two fractions.
**Which fraction is closer to \sqrt{2**?} We compare the absolute distances:
- Distance 1:
- Distance 2:
To see which distance is smaller, we compare with . This is equivalent to comparing with , which is equivalent to comparing with , or with . Since are positive integers, , so . This implies . Therefore, Distance 1 is always greater than Distance 2. The new fraction is always closer to .
Visualization:
[TikZ diagram omitted in web version]
This visualization shows how the process of generating new fractions from always produces a better approximation to . The red dashed line represents , and the arrows show the improvement in approximation.
∎
Problem 1.14: Irrationality of
Section titled “Problem 1.14: Irrationality of n−1+n+1\sqrt{n - 1} + \sqrt{n + 1}n−1+n+1”Problem. Prove that is irrational for every integer .
Strategy: Use proof by contradiction. Assume the sum is rational, then square both sides to eliminate the square roots. This leads to being rational, which is false since is not a perfect square for .
Solution: Assume , where are integers, , .
Square both sides:
Thus:
Suppose is rational, say , . Then:
For , , irrational. For , is not a perfect square (since ). If , must be a perfect square, a contradiction for . Thus, is irrational, so is irrational.
∎
Problem 1.15: Approximation by Rational Numbers
Section titled “Problem 1.15: Approximation by Rational Numbers”Problem. Given a real and an integer , prove that there exist integers and with such that . Hint. Consider the numbers for and show that some pair differs by at most .
Strategy: Use the pigeonhole principle. Consider the fractional parts of , divide into equal subintervals, and apply the pigeonhole principle to find two numbers in the same subinterval.
Solution: We will use the pigeonhole principle to prove this result.
Consider the numbers: .
Let’s look at the fractional parts of these numbers. The fractional part of a number is , where is the greatest integer less than or equal to .
The fractional parts of all lie in the interval .
Divide the interval into equal subintervals:
By the pigeonhole principle, since we have numbers and only subintervals, at least two of these numbers must fall into the same subinterval.
Let’s say and (where ) have fractional parts in the same subinterval. Then:
Let and . Then:
Since , we have , and is an integer.
Therefore, we have found integers and with such that .
Example: If and , we might find that and have fractional parts in the same subinterval, giving us .
∎
Problem 1.16: Infinitely Many Rational Approximations
Section titled “Problem 1.16: Infinitely Many Rational Approximations”Problem. If is irrational, prove that there are infinitely many rational numbers with such that .
Strategy: We will use the result from Problem 1.15 (Dirichlet’s Approximation Theorem) to construct rational approximations, then use proof by contradiction to show that there must be infinitely many distinct such approximations.
Solution: We will construct an infinite sequence of distinct rational numbers satisfying the condition.
From Problem 1.15 (Dirichlet’s Approximation Theorem), for any integer , there exist integers and with such that . Dividing by , we get:
Since , we have , which implies . Thus, for any integer , we can find a rational number such that:
Now we must show that this process can generate infinitely many distinct fractions. Assume, for the sake of contradiction, that there are only a finite number of such rational approximations, say . Since is irrational, for any rational number , the distance is non-zero. Let be the smallest of these non-zero distances:
Now, choose an integer large enough such that . By the result from Problem 1.15, there exist integers and with such that:
This implies . So we have found a new rational approximation such that:
Since the approximation error of is smaller than , cannot be one of the fractions in our finite list . This contradicts our assumption that we had a complete list of all such approximations. Therefore, there must be infinitely many such rational numbers.
∎
Problem 1.17: Factorial Representation of Rationals (Precise Form)
Section titled “Problem 1.17: Factorial Representation of Rationals (Precise Form)”Problem. Let be a positive rational number of the form
where each is a nonnegative integer with for and . Let denote the greatest integer less than or equal to . Prove that , that
and that is the smallest integer such that is an integer. Conversely, show that every positive rational number can be expressed in this form in one and only one way.
Strategy: For the forward direction, use the properties of the factorial series to show that the fractional part is less than 1, then use the floor function properties. For the converse, use proof by contradiction to show uniqueness by assuming two different representations and finding a contradiction.
Solution: Let with the given conditions on .
1. Proof that : The sum can be written as . We must show the summation part is a positive fraction less than 1. Since , the sum is positive. We can bound the sum using the property :
The infinite sum is a known identity: . This is a telescoping series whose sum is the first term, . Thus, . This means , so by definition, .
2. Formula for : Define . Then is an integer for . Consider the expression this gets complicated. Let’s use the given formula. Let . From part (1), we know . So . Let’s test the formula: . We have and . So, . This proves the formula for .
3. Minimality of : Multiplying by gives . Since , each term is an integer, so is an integer. For any , when we compute , the term corresponding to is . Since , this term is a non-integer fraction. Because all other terms for are also fractions and terms for are integers, cannot be an integer. Thus, is the smallest such integer.
4. Converse (Uniqueness): Suppose a positive rational number has two different representations:
with the conditions for , , and similarly for . From part (3), is the smallest integer such that is an integer, and is the smallest integer such that is an integer. This implies .
Let be the largest index for which the coefficients differ, so . Assume, without loss of generality, that . Since for , we can subtract the sums:
Rearranging the terms, we get:
Multiply both sides by :
The right-hand side is an integer, because for each , divides . Let’s analyze the left-hand side. Since and are integers and , we have . From the conditions on the coefficients, (for ) and . Therefore, . This implies that for , the left-hand side is a non-integer fraction, since the numerator is an integer between and , and the denominator is . This creates a contradiction: the left-hand side cannot be an integer, while the right-hand side must be an integer. For the case , the equation becomes , which contradicts . Thus, our assumption that there is a largest index where must be false. All coefficients must be identical. The representation is unique.
5. Uniqueness: Suppose has two different representations, . Let be the largest index where . Assume . Then . The left side is . The right side is bounded above by , a contradiction. Thus, all coefficients must be the same.
∎
Upper Bounds
Section titled “Upper Bounds”Definitions and Theorems
Section titled “Definitions and Theorems”Theorem (Completeness Axiom).
Every nonempty set of real numbers that is bounded above has a supremum.
**Importance:**The Completeness Axiom is the fundamental property that distinguishes the real numbers from the rational numbers. It ensures that the real number system has no “gaps” and is essential for all of calculus and analysis. This axiom is the foundation for proving the existence of limits, continuity, and many other important results in mathematics.
Theorem (Uniqueness of Supremum and Infimum).
If a set has a supremum (infimum), it is unique.
**Importance:**This theorem ensures that suprema and infima are well-defined concepts. Without uniqueness, we couldn’t meaningfully talk about “the” supremum or “the” infimum of a set. This result is essential for the logical consistency of real analysis and is used implicitly in many proofs throughout mathematics.
Theorem (Archimedean Property).
For any positive real numbers and , there exists a positive integer such that .
Theorem (Comparison Property for Suprema).
Let and be nonempty subsets of such that for all and . If has a supremum, then has a supremum and .
**Importance:**This theorem provides a powerful tool for comparing suprema of different sets. It’s essential for many proofs in analysis where we need to establish inequalities between suprema. This result is particularly useful in functional analysis, measure theory, and optimization theory where we often need to compare bounds of different sets or functions.
Problem 1.18: Uniqueness of Supremum and Infimum
Section titled “Problem 1.18: Uniqueness of Supremum and Infimum”Problem. Show that the and of a set are uniquely determined whenever they exist.
Strategy: We will prove the uniqueness of supremum by contradiction, assuming there are two different suprema and showing this leads to a contradiction. The same approach applies to infimum.
Solution: We will prove that if a set has a supremum, it is unique. The proof for infimum is similar.
Proof by contradiction: Suppose a set has two different suprema, say and , with .
Since is a supremum of :
- is an upper bound of (every element of is )
- is the least upper bound (no number less than is an upper bound)
Since is also a supremum of :
- is an upper bound of (every element of is )
- is the least upper bound (no number less than is an upper bound)
But since , the number is less than and is also an upper bound of . This contradicts the fact that is the least upper bound.
Therefore, our assumption that there are two different suprema is false, and the supremum must be unique.
Alternative proof: Let and both be suprema of . Then:
- is an upper bound, so (since is the least upper bound)
- is an upper bound, so (since is the least upper bound)
Therefore, .
For infimum: The same argument applies to infimum. If a set has two infima and , then:
- is a lower bound, so (since is the greatest lower bound)
- is a lower bound, so (since is the greatest lower bound)
Therefore, .
∎
Problem 1.19: Finding Supremum and Infimum
Section titled “Problem 1.19: Finding Supremum and Infimum”Problem. Find the and of each of the following sets:
- All numbers of the form for positive integers .
- .
- where .
Strategy: For (a), analyze the range of each exponential term and find the maximum and minimum values. For (b), solve the quadratic inequality to find the interval. For (c), use the sign changes of the polynomial at its roots to determine the intervals where the product is negative.
Solution:
**1. Numbers of the form 2^{-p** + 3^{-q} + 5^{-r}:}
Let’s analyze the range of each term:
- ranges from (when ) to (as )
- ranges from (when ) to (as )
- ranges from (when ) to (as )
Therefore:
- The maximum value occurs when :
- The minimum value occurs as :
So \frac{31**{30}} and 0.
**2. Set :}
First, let’s find the roots of :
Since the coefficient of is positive, the parabola opens upward. The inequality holds between the roots.
Therefore, , so 3 and \frac{1**{3}}.
Visualization for part (b):
[TikZ diagram omitted in web version]
This visualization shows the quadratic function and highlights the interval where the inequality holds, which is between the roots and .
**3. Set where :}
The expression changes sign at each root .
Starting from :
- For : all factors are negative, so the product is positive
- For : one factor is positive, three negative, so product is negative
- For : two factors positive, two negative, so product is positive
- For : three factors positive, one negative, so product is negative
- For : all factors are positive, so product is positive
Therefore, .
So d and a.
Visualization for part (c):
[TikZ diagram omitted in web version]
This visualization shows the fourth-degree polynomial and highlights the intervals where the inequality holds. The polynomial changes sign at each root, creating alternating positive and negative intervals.
∎
Problem 1.20: Comparison Property for Suprema
Section titled “Problem 1.20: Comparison Property for Suprema”Problem. Let ( S ) and ( T ) be nonempty subsets of ( \mathbb{R} ) such that ( s \leq t ) for all ( s \in S ) and ( t \in T ). Suppose ( T ) has a supremum. Then ( S ) has a supremum and
Strategy: We will first show that has a supremum by using the completeness axiom, then prove that by showing that is a lower bound for and using the definition of supremum.
Solution: Let and be nonempty subsets of with the property that for every and , we have .
-
Existence of : Since is nonempty, we can pick an arbitrary element . By the given property, for every , we have . This shows that is bounded above (by any element of ). Since is also nonempty and bounded above, the completeness axiom of guarantees that exists. Let’s call it .
-
Proof that : Let and . From step 1, we know that any element is an upper bound for the set . Since is the least upper bound of , it must be less than or equal to any other upper bound of . Therefore, for any , we must have:
This inequality shows that is a lower bound for the set . Now, by definition, is the least upper bound of . As an upper bound for , must be greater than or equal to every element of . More importantly, it must be greater than or equal to any lower bound of . Since we have established that is a lower bound for , it must follow that:
Substituting the definitions of and , we get:
This completes the proof.
∎
Problem 1.21: Product of Suprema
Section titled “Problem 1.21: Product of Suprema”Problem. Let ( A ) and ( B ) be two sets of positive real numbers, each bounded above. Let ( a = \sup A ), ( b = \sup B ). Define
Prove that
Strategy: We will show that is an upper bound for , then prove it is the least upper bound by using the definition of supremum and constructing elements of that are arbitrarily close to .
Solution:
Since ( A ) and ( B ) are sets of positive real numbers bounded above, their suprema ( a = \sup A ) and ( b = \sup B ) exist and are finite.
We are to prove that:
Step 1: Show that ( ab ) is an upper bound for ( C ).
Let ( x \in A ), ( y \in B ). Since ( x \leq a ) and ( y \leq b ), we have:
Therefore, every element ( c \in C ) satisfies ( c \leq ab ), so ( ab ) is an upper bound for ( C ).
Step 2: Show that ( ab ) is the least upper bound.
Let ( \varepsilon > 0 ). Since ( a = \sup A ), there exists ( x_\varepsilon \in A ) such that:
Similarly, since ( b = \sup B ), there exists ( y_\varepsilon \in B ) such that:
Now consider:
Since ( \frac{\varepsilon^2}{4ab} > 0 ), we have:
Therefore, for every ( \varepsilon > 0 ), there exists ( c \in C ) such that ( c > ab - \varepsilon ). Hence, ( ab ) is the least upper bound of ( C ).
∎
Problem 1.22: Representation of Rationals in Base ( k )
Section titled “Problem 1.22: Representation of Rationals in Base ( k )”Problem. Given ( x \geq 0 ) and an integer ( k \geq 2 ), let ( a_0 ) denote the largest integer ( \leq x ), and, assuming that ( a_0, a_1, \dots, a_{n-1} ) have been defined, let ( a_n ) denote the largest integer such that
- [(a)] Prove that ( 0 \leq a_i \leq k - 1 ) for each ( i = 1, 2, \dots ).
- [(b)] Let ( r_n = a_0 + a_1 k^{-1} + a_2 k^{-2} + \cdots + a_n k^{-n} ) and show that ( x = \sup { r_n } ), the supremum of the set of rational numbers ( r_1, r_2, \dots ).
Strategy: For (a), use the definition of as the largest integer satisfying the condition and show that choosing would violate it. For (b), use the Archimedean property and proof by contradiction to show that the supremum equals .
Solution: Let ( r_n = \sum_{i=0}^n \frac{a_i}{k^i} ). By definition, ( a_n ) is the largest integer such that ( r_n \leq x ).
(a) Show ( 0 \leq a_i \leq k - 1 ): Since ( a_n ) is the largest integer satisfying the condition, choosing ( a_n + 1 ) would violate it:
From the definition of (a_{n-1}), we know it was the largest integer such that ( r_{n-1} \le x ). This implies . Now, from the definition of , we have , which implies . Combining these facts:
Since is an integer and , we must have . Also, must be non-negative, otherwise we could choose to get a larger (or equal) sum that is still less than or equal to , contradicting the “largest integer” definition if the original were negative. Thus, .
*(b) Show that ( x = \sup { r_n * ):} The sequence is non-decreasing by construction, since . It is also bounded above by . Therefore, its supremum exists; let . We know . We will prove by contradiction. Assume . Let . By the Archimedean property, we can choose an integer large enough such that . From the definition of , we know and . The second inequality rearranges to . Since , we know . Therefore, . Substituting , we get . But we chose such that . This gives , a contradiction. Thus, our assumption must be false, and .
∎
Inequalities and Identities
Section titled “Inequalities and Identities”Additional Theorems for Inequalities and Identities
Section titled “Additional Theorems for Inequalities and Identities”Theorem (Binomial Theorem).
For any real numbers and and positive integer :
**Importance:**The Binomial Theorem is one of the most fundamental results in algebra and combinatorics. It provides a formula for expanding powers of binomials and is essential for understanding polynomials, probability theory, and many areas of mathematics. The binomial coefficients that appear in this theorem are fundamental in combinatorics and have applications throughout mathematics.
Theorem (Pigeonhole Principle).
If objects are placed into containers where , then at least one container must contain more than one object.
**Importance:**The Pigeonhole Principle is a simple but powerful tool in combinatorics and discrete mathematics. It provides a way to prove the existence of certain patterns or properties without explicitly constructing them. This principle is essential for many existence proofs and has applications in computer science, number theory, and many other areas.
Theorem (Dirichlet’s Approximation Theorem).
For any real number and positive integer , there exist integers and with such that .
**Importance:**Dirichlet’s Approximation Theorem is a fundamental result in Diophantine approximation that shows how well real numbers can be approximated by rational numbers. It’s essential for understanding the distribution of rational numbers and has applications in number theory, analysis, and many areas of mathematics. This theorem is the foundation for more advanced results in Diophantine approximation.
Problem 1.23: Lagrange’s Identity
Section titled “Problem 1.23: Lagrange’s Identity”Problem. Prove Lagrange’s identity for real numbers:
Strategy: We will prove this identity by expanding both sides and showing they are equal. We’ll expand the left-hand side as a double sum and the right-hand side by expanding the product and the squared terms, then show that all terms cancel appropriately.
Solution: We will prove Lagrange’s identity by expanding both sides and showing they are equal.
Let’s start by expanding the left-hand side:
Now let’s expand the right-hand side:
Let’s simplify this step by step. First, note that:
Substituting this back into our expression:
This is exactly the same as our expanded left-hand side! Therefore, Lagrange’s identity is proven.
Alternative Proof using Determinants: We can also prove this using the fact that the determinant of a 2×2 matrix is zero if and only if its rows are linearly dependent.
Consider the matrix:
The determinant of this matrix is . If we square this determinant and sum over all pairs with , we get the right-hand side of Lagrange’s identity.
The left-hand side represents the square of the dot product of the vectors and .
The identity shows that the square of the dot product equals the product of the squared magnitudes minus the sum of squared determinants of all 2×2 submatrices formed by pairs of components.
∎
Problem 1.24: A Holder-type Inequality
Section titled “Problem 1.24: A Holder-type Inequality”Problem. Prove that for arbitrary real numbers ( a_k, b_k, c_k ) we have
Strategy: We will apply the Cauchy-Schwarz inequality twice. First, we’ll group and , then apply Cauchy-Schwarz to the resulting term by treating it as the dot product of sequences and .
Solution: We will prove this inequality by applying the Cauchy-Schwarz inequality twice. First, group the terms as and . Applying the Cauchy-Schwarz inequality to the sequences and gives:
Next, we apply the Cauchy-Schwarz inequality to the term , treating it as the dot product of sequences and :
This implies:
Now, substitute this result back into our first inequality:
Finally, squaring both sides gives the desired result:
∎
Problem 1.25: Minkowski’s Inequality
Section titled “Problem 1.25: Minkowski’s Inequality”Problem. Prove Minkowski’s inequality:
Strategy: We will expand the left-hand side, apply the Cauchy-Schwarz inequality to the cross term, then complete the square to obtain the desired inequality.
Solution:
Let ( A = \left( \sum a_k^2 \right)^{1/2} ), ( B = \left( \sum b_k^2 \right)^{1/2} ), and expand the square:
Apply Cauchy–Schwarz:
Thus,
Taking square roots:
∎
Problem 1.26: Chebyshev’s Sum Inequality
Section titled “Problem 1.26: Chebyshev’s Sum Inequality”Problem. If ( a_1 \geq a_2 \geq \cdots \geq a_n ) and ( b_1 \geq b_2 \geq \cdots \geq b_n ), prove that
Strategy: We will consider the double summation and show that it is non-negative due to the ordering of the sequences, then expand it to obtain the desired inequality.
Solution: Consider the double summation
Since the sequences and are sorted in the same order (both non-increasing), the terms and always have the same sign. If , then and , so both differences are non-positive. If , both are non-negative. Therefore, their product is always non-negative:
This implies that the total sum must be non-negative, .
Now, let’s expand the sum:
We evaluate each double summation:
- ( \sum_{i=1}^n \sum_{j=1}^n a_i b_i = \sum_{i=1}^n \left( n \cdot a_i b_i \right) = n \sum_{i=1}^n a_i b_i )
- ( \sum_{i=1}^n \sum_{j=1}^n a_i b_j = \left( \sum_{i=1}^n a_i \right) \left( \sum_{j=1}^n b_j \right) )
- ( \sum_{i=1}^n \sum_{j=1}^n a_j b_i = \left( \sum_{j=1}^n a_j \right) \left( \sum_{i=1}^n b_i \right) )
- ( \sum_{i=1}^n \sum_{j=1}^n a_j b_j = \sum_{j=1}^n \left( n \cdot a_j b_j \right) = n \sum_{j=1}^n a_j b_j )
Substituting these back into the expression for :
Since we established that :
Dividing by 2 and rearranging gives the desired inequality:
∎
Complex Numbers
Section titled “Complex Numbers”Problem 1.27: Express Complex Numbers in Form
Section titled “Problem 1.27: Express Complex Numbers in a+bia + bia+bi Form”Problem. Express the following complex numbers in the form ( a + bi ):
- [(a)] ( (1 + i)^3 )
- [(b)] ( \frac{2 + 3i}{3 - 4i} )
- [(c)] ( i^5 + i^{16} )
- [(d)] ( \frac{1}{2}(1 + i)(1 + i^{-8}) )
Strategy: We will use the properties of complex numbers, including , , and the fact that powers of cycle every 4. For division, we’ll rationalize the denominator by multiplying by the complex conjugate.
Solution:
- [(a)] ( (1 + i)^3 = (1 + i)^2 (1 + i) = (2i)(1 + i) = 2i + 2i^2 = 2i - 2 = -2 + 2i )
- [(b)] Rationalize the denominator:
- [(c)] ( i^5 = i ), since ( i^4 = 1 ), and ( i^{16} = (i^4)^4 = 1 ), so:
- [(d)] ( \frac{1}{2}(1 + i)(1 + i^{-8}) ), note that ( i^{-8} = (i^4)^{-2} = 1^{-2} = 1 ), so:
∎
Problem 1.28: Solve Complex Equations
Section titled “Problem 1.28: Solve Complex Equations”Problem. In each case, determine all real ( x ) and ( y ) which satisfy the given relation:
- [(a)] ( x + iy = |x - iy| )
- [(b)] ( x + iy = (x - iy)^2 )
- [(c)] ( \sum_{k=0}^{100} i^k = x + iy )
Strategy: For each equation, we’ll equate the real and imaginary parts. For (a), we’ll use the fact that the right-hand side is real and nonnegative. For (b), we’ll expand the square and solve the resulting system. For (c), we’ll use the cyclic nature of powers of .
Solution:
- [(a)] RHS is real and nonnegative. LHS is complex. For equality, imaginary part must vanish:
So solution: ( y = 0,, x \geq 0 )
- [(b)] Compute RHS:
Set equal to ( x + iy ), equate real and imaginary parts:
From second equation: ( y = -2xy \Rightarrow y(1 + 2x) = 0 \Rightarrow y = 0 ) or ( x = -\frac{1}{2} )
If ( y = 0 ), then first equation: ( x = x^2 \Rightarrow x(x - 1) = 0 \Rightarrow x = 0 ) or ( x = 1 )
If ( x = -\frac{1}{2} ), then first equation:
So all solutions:
- [(c)] The powers of ( i ) cycle every 4: ( i^0 = 1, i^1 = i, i^2 = -1, i^3 = -i )
There are ( 101 ) terms, which form 25 full cycles and one leftover term ( i^{100} \equiv i^0 = 1 )
Each full cycle sums to 0. So total sum:
∎
Problem 1.29: Basic Identities for Complex Conjugates
Section titled “Problem 1.29: Basic Identities for Complex Conjugates”Problem. If ( z = x + iy ), where ( x ) and ( y ) are real, the complex conjugate of ( z ) is ( \overline{z} = x - iy ). Prove the following:
- ( \overline{z_1 + z_2} = \overline{z_1} + \overline{z_2} ),
- ( \overline{z_1 z_2} = \overline{z_1} \cdot \overline{z_2} ),
- ( z \cdot \overline{z} = |z|^2 ),
- ( z + \overline{z} ) is twice the real part of ( z ),
- ( \frac{z - \overline{z}}{i} ) is twice the imaginary part of ( z ).
Strategy: We will prove each identity by using the definition of complex conjugate and performing the necessary algebraic manipulations. For each part, we’ll work with the real and imaginary components explicitly.
Solution: Let ( z = x + iy ) and ( w = u + iv ) be two complex numbers.
- Conjugate of a sum:
- Conjugate of a product:
- Modulus squared:
- Twice the real part:
- Twice the imaginary part:
∎
Problem 1.30: Geometric Descriptions of Complex Sets
Section titled “Problem 1.30: Geometric Descriptions of Complex Sets”Problem. Describe geometrically the set of complex numbers ( z ) which satisfies each of the following conditions:
- ( |z| = 1 ),
- ( |z| < 1 ),
- ( |z| \leq 1 ),
- ( z + \overline{z} = 1 ),
- ( z - \overline{z} = i ),
- ( \overline{z} + z = |z|^2 ).
Strategy: We will use the properties of complex conjugates and the relationship between complex numbers and their real/imaginary parts to translate each condition into geometric terms. For the last condition, we’ll complete the square to identify the geometric shape.
Solution:
- The unit circle centered at the origin.
- The open unit disk centered at the origin.
- The closed unit disk centered at the origin.
- ( 2 \Re(z) = 1 \Rightarrow \Re(z) = \frac{1}{2} ): a vertical line in the complex plane.
- ( 2i \Im(z) = i \Rightarrow \Im(z) = \frac{1}{2} ): a horizontal line.
- Let ( z = x + iy ), where ( x, y \in \mathbb{R} ). Then:
So the equation becomes:
Rewriting this:
We now complete the square on the ( x )-terms:
which gives:
This is the standard equation of a circle with center at ( (1, 0) ) and radius ( 1 ) in the complex plane.
Visualizations:
[Table omitted in web version]
∎
Problem 1.31: Equilateral Triangle on the Unit Circle
Section titled “Problem 1.31: Equilateral Triangle on the Unit Circle”Problem. Given three complex numbers ( z_1, z_2, z_3 ) such that ( |z_1| = |z_2| = |z_3| = 1 ) and ( z_1 + z_2 + z_3 = 0 ), show that these numbers are vertices of an equilateral triangle inscribed in the unit circle with center at the origin.
Strategy: Use the fact that the sum of three unit complex numbers equals zero to show they must be the cube roots of unity (rotated), which form an equilateral triangle. Verify that the angles differ by and the sum condition is satisfied.
Solution: Since ( |z_i| = 1 ), each ( z_i = e^{i\theta_i} ) lies on the unit circle. Given ( z_1 + z_2 + z_3 = 0 ), we need to show they form an equilateral triangle. Consider the angles ( \theta_1, \theta_2, \theta_3 ). The sum condition implies:
For three points on the unit circle to form an equilateral triangle, their arguments must differ by ( 120^\circ = \frac{2\pi}{3} ). Assume:
Check the sum:
Since ( e^{i\frac{2\pi}{3}} = -\frac{1}{2} + i\frac{\sqrt{3}}{2} ), ( e^{i\frac{4\pi}{3}} = -\frac{1}{2} - i\frac{\sqrt{3}}{2} ), we have:
The angles ( \theta, \theta + \frac{2\pi}{3}, \theta + \frac{4\pi}{3} ) are spaced ( \frac{2\pi}{3} ) apart, forming an equilateral triangle. Any three points with ( |z_i| = 1 ) and sum zero are rotations of the cube roots of unity, ensuring an equilateral triangle.
Visualization:
[TikZ diagram omitted in web version]
This visualization shows the equilateral triangle formed by the cube roots of unity on the unit circle. The green vectors show how the sum of the three complex numbers equals zero, and the orange angle markers show the spacing between vertices.
∎
Problem 1.32: Inequality with Complex Numbers
Section titled “Problem 1.32: Inequality with Complex Numbers”Problem. If ( a ) and ( b ) are complex numbers, prove:
- ( |a - b|^2 \leq (1 + |a|^2)(1 + |b|^2) ),
- If ( a \neq 0 ), then ( |a + b| = |a| + |b| ) if and only if ( \frac{b}{a} ) is real and nonnegative.
Strategy: For part (a), we’ll expand both sides and show that the difference is non-negative. For part (b), we’ll use the fact that equality in the triangle inequality occurs when the complex numbers are collinear and point in the same direction.
Solution:
- Compute:
Consider the right-hand side:
Evaluate:
Since ( |a b|^2 \geq 0 ), ( \Re(a\overline{b}) \geq -|a b| ):
Thus, ( |a - b|^2 \leq (1 + |a|^2)(1 + |b|^2) ).
- For ( |a + b| = |a| + |b| ), the triangle inequality requires ( a, b ) collinear in the same direction. Let ( b = ka ), ( k \in \mathbb{R}_{\geq 0} ):
Thus, ( \frac{b}{a} = k \geq 0 ). Conversely, if ( |a + b| = |a| + |b| ), then ( a\overline{b} + \overline{a}b = 2|a||b| ), so ( \frac{b}{a} ) is real and nonnegative.
∎
Problem 1.33: Equality Condition for Complex Difference
Section titled “Problem 1.33: Equality Condition for Complex Difference”Problem. If ( a ) and ( b ) are complex numbers, prove that
if and only if ( |a| = 1 ) or ( |b| = 1 ). For which ( a ) and ( b ) is the inequality ( |a - b| < |1 - \overline{a}b| ) valid?
Strategy: We will compute the difference and show that it factors as where and . This will allow us to determine when equality holds and when the inequality is valid.
Solution: Let ( |a| = r ), ( |b| = s ). Compute:
Thus:
Equality holds when:
For the inequality:
Thus, equality holds if ( |a| = 1 ) or ( |b| = 1 ); the inequality holds when one modulus is less than 1 and the other is greater than 1.
∎
Problem 1.34: Complex Circle in the Plane
Section titled “Problem 1.34: Complex Circle in the Plane”Problem. If ( a ) and ( c ) are real constants, ( b ) is complex, show that the equation
represents a circle in the ( xy )-plane.
Strategy: We will substitute and into the equation, then use the fact that and to show that the equation reduces to the general form of a circle.
Solution: Let ( z = x + iy ), ( \overline{z} = x - iy ), then ( z \overline{z} = x^2 + y^2 ), ( bz + \overline{b} \overline{z} = 2 \Re(b z) ). Hence the equation becomes:
This is the general form of a circle in ( \mathbb{R}^2 ).
∎
Problem 1.35: Argument of a Complex Number via Arctangent
Section titled “Problem 1.35: Argument of a Complex Number via Arctangent”Problem. Recall the definition of the inverse tangent: given a real number ( t ), ( \tan^{-1}(t) ) is the unique real number ( \theta ) satisfying:
If ( z = x + iy ), show that:
- ( \arg(z) = \tan^{-1}\left( \frac{y}{x} \right) ), if ( x > 0 ),
- ( \arg(z) = \tan^{-1}\left( \frac{y}{x} \right) + \pi ), if ( x < 0 ), ( y \geq 0 ),
- ( \arg(z) = \tan^{-1}\left( \frac{y}{x} \right) - \pi ), if ( x < 0 ), ( y < 0 ),
- ( \arg(z) = \frac{\pi}{2} ), if ( x = 0, y > 0 ); \quad ( \arg(z) = -\frac{\pi}{2} ), if ( x = 0, y < 0 ).
Strategy: We will use the relationship between the argument of a complex number and the quadrant it lies in. The principal value of gives angles in , so we need to adjust for different quadrants to get the correct argument in .
Solution: For ( z = x + iy ), ( \arg(z) ) is the angle ( \theta \in (-\pi, \pi] ) such that ( z = |z| e^{i\theta} ).
- If ( x > 0 ), ( z ) is in Quadrant I or IV, and ( \tan \theta = \frac{y}{x} ), so ( \theta = \tan^{-1}\left( \frac{y}{x} \right) ).
- If ( x < 0 ), ( y \geq 0 ), ( z ) is in Quadrant II. ( \tan^{-1}\left( \frac{y}{x} \right) \in (-\frac{\pi}{2}, 0] ), so add ( \pi ) to get ( \theta \in (\frac{\pi}{2}, \pi] ).
- If ( x < 0 ), ( y < 0 ), ( z ) is in Quadrant III. ( \tan^{-1}\left( \frac{y}{x} \right) \in (0, \frac{\pi}{2}] ), so subtract ( \pi ) to get ( \theta \in (-\pi, -\frac{\pi}{2}] ).
- If ( x = 0 ), ( z = iy ). If ( y > 0 ), ( \theta = \frac{\pi}{2} ); if ( y < 0 ), ( \theta = -\frac{\pi}{2} ).
∎
Problem 1.36: Pseudo-Ordering on Complex Numbers
Section titled “Problem 1.36: Pseudo-Ordering on Complex Numbers”Problem. Define the following pseudo-ordering on complex numbers: if , or if and . Which of Axioms 6,7,8,9 are satisfied by this relation?
Strategy: We will examine each axiom individually, testing whether the pseudo-ordering satisfies the properties of trichotomy, translation invariance, multiplication invariance, and transitivity. We’ll provide counterexamples where axioms fail.
-
Axiom 6 (Trichotomy): For any , we can compare their moduli. Exactly one of , , or holds. If , we compare their principal arguments, for which trichotomy holds on . Thus, exactly one of , , or is true. This axiom is satisfied.
-
Axiom 9 (Transitivity): If and , the transitivity of the relation on the real numbers for both the moduli and the arguments ensures that . This axiom is satisfied.
-
Axiom 7 (Translation Invariance): This axiom states that if , then for any . This axiom is not satisfied.
Counterexample: Let and . According to the ordering, because . Now, let . Then . And . We must compare and . We have and . Since , we have in this pseudo-ordering. So, . The order relation was reversed, which violates the axiom.
-
Axiom 8 (Multiplication): This axiom states that if and , then . Let us define to mean . This holds for any . This axiom is also not satisfied.
Counterexample: Let and . We have . The arguments are and . Since , we have . Now, let . Since , is a “positive” number under this definition. Then . And . We must compare and . We have and . The arguments are and . Since , we have . So, . The order relation was reversed from to . The axiom is violated.
Conclusion: Axioms 6 and 9 are satisfied; Axiom 7 and 8 is not applicable.
Solution: The pseudo-ordering on complex numbers is defined by if either or and .
Axiom 6 (Trichotomy): For any , exactly one of the following holds:
- , in which case
- , in which case
- , in which case we compare arguments: \begin{enumerate}[label=(\roman*)]
- , so
- , so
- , so
\end{enumerate} Therefore, Axiom 6 is satisfied.
Axiom 7 (Translation Invariance): This axiom is not satisfied. Consider and . We have since . However, with , we get and . Since , we have , reversing the order.
Axiom 8 (Multiplication Invariance): This axiom is not satisfied. Consider and . Since and , we have . However, with , we get and . Since and , we have , again reversing the order.
Axiom 9 (Transitivity): If and , then either:
- , so and
- , so and
- , so and
- and , so and
Therefore, Axiom 9 is satisfied.
In summary, Axioms 6 and 9 are satisfied, while Axioms 7 and 8 are not satisfied.
∎
Problem 1.37: Order Axioms and Lexicographic Ordering on
Section titled “Problem 1.37: Order Axioms and Lexicographic Ordering on R2\mathbb{R}^2R2”Problem. Define a pseudo-ordering on ordered pairs ((x_1, y_1) < (x_2, y_2)) if either
- (x_1 < x_2), or
- (x_1 = x_2) and (y_1 < y_2).
Which of Axioms 6, 7, 8, 9 are satisfied by this relation?
Strategy: We will examine each axiom for the lexicographic ordering on . This ordering compares first coordinates, then second coordinates if the first coordinates are equal, which should preserve most of the standard ordering properties.
Solution:
- Axiom 6: Trichotomy. For any ( (x_1, y_1), (x_2, y_2) ), if ( x_1 < x_2 ), then ( (x_1, y_1) < (x_2, y_2) ); if ( x_1 > x_2 ), then ( (x_2, y_2) < (x_1, y_1) ); if ( x_1 = x_2 ), compare ( y_1, y_2 ). Exactly one holds. Satisfied.
- Axiom 7: Translation Invariance. If ( (x_1, y_1) < (x_2, y_2) ), add ( (u, v) ): if ( x_1 < x_2 ), then ( x_1 + u < x_2 + u ); if ( x_1 = x_2 ), then ( y_1 < y_2 \implies y_1 + v < y_2 + v ). Satisfied.
- Axiom 8: Multiplication. Not applicable, as ( \mathbb{R}^2 ) lacks scalar multiplication.
- Axiom 9: Transitivity. If ( (x_1, y_1) < (x_2, y_2) ), ( (x_2, y_2) < (x_3, y_3) ), lexicographic order ensures ( (x_1, y_1) < (x_3, y_3) ). Satisfied.
Conclusion: Axioms 6, 7, and 9 are satisfied; Axiom 8 is not applicable.
∎
Problem 1.38: Argument of a Quotient Using Theorem 1.48
Section titled “Problem 1.38: Argument of a Quotient Using Theorem 1.48”Problem. State and prove a theorem analogous to Theorem 1.48, expressing ( \arg\left( \frac{z_1}{z_2} \right) ) in terms of ( \arg(z_1) ) and ( \arg(z_2) ).
Strategy: We will use the fact that and apply Theorem 1.48 to the product, using the property that .
Solution: Theorem: If ( z_1, z_2 \neq 0 ), then:
where ( n(z_1, z_2^{-1}) ) adjusts the argument to ( (-\pi, \pi] ).
Solution: Since ( \frac{z_1}{z_2} = z_1 z_2^{-1} ), and ( \arg(z_2^{-1}) = -\arg(z_2) ), apply Theorem 1.48:
∎
Problem 1.39: Logarithm of a Quotient Using Theorem 1.54
Section titled “Problem 1.39: Logarithm of a Quotient Using Theorem 1.54”Problem. State and prove a theorem analogous to Theorem 1.54, expressing ( \log\left( \frac{z_1}{z_2} \right) ) in terms of ( \log(z_1) ) and ( \log(z_2) ).
Strategy: We will use the fact that and apply Theorem 1.54 to the product, using the property that .
Solution: Theorem: If ( z_1, z_2 \neq 0 ), then:
Solution: Since ( \frac{z_1}{z_2} = z_1 z_2^{-1} ), apply Theorem 1.54:
∎
Problem 1.40: Roots of Unity and Polynomial Identity
Section titled “Problem 1.40: Roots of Unity and Polynomial Identity”Problem. Prove that the ( n )th roots of 1 are given by ( \alpha, \alpha^2, \ldots, \alpha^n ), where ( \alpha = e^{2\pi i/n} ), and that these roots ( \ne 1 ) satisfy the equation
Strategy: We will use the fact that the th roots of unity are the solutions to , and use the factorization to show that all roots except satisfy the given equation.
Solution: Let ( \alpha = e^{2\pi i/n} ). Then ( \alpha^n = 1 ), so it’s a root of ( x^n - 1 = 0 ). Also,
∎
Problem 1.41: Inequalities and Boundedness of cos z
Section titled “Problem 1.41: Inequalities and Boundedness of cos z”Problem.
- Prove that ( |z^i| < e^{\pi} ) for all complex ( z \ne 0 ).
- Prove that there is no constant ( M > 0 ) such that ( |\cos z| < M ) for all complex ( z ).
Strategy: For part (a), we’ll use the definition and analyze the modulus in terms of the argument. For part (b), we’ll use the fact that which grows exponentially as .
Solution:
- For ( z = re^{i\theta} ), ( z^i = e^{i(\ln r + i\theta)} = e^{-\theta} e^{i \ln r} ), so ( |z^i| = e^{-\theta} ). Since ( \theta \in (-\pi, \pi] ), ( |z^i| \leq e^{\pi} ), strict unless ( \theta = -\pi ).
- For ( z = iy ), ( \cos(iy) = \cosh y ), which is unbounded as ( |y| \to \infty ). Thus, no ( M > 0 ) exists.
∎
Problem 1.42: Complex Exponential via Real and Imaginary Parts
Section titled “Problem 1.42: Complex Exponential via Real and Imaginary Parts”Problem. If ( w = u + iv ), where ( u ) and ( v ) are real, show that
Strategy: We will use the definition and expand the product to separate the real and imaginary parts.
Solution: For ( z^w = e^{w \log z} ), where ( \log z = \log |z| + i \arg z ):
Thus:
∎
Problem 1.43: Logarithmic Identities for Complex Powers
Section titled “Problem 1.43: Logarithmic Identities for Complex Powers”Problem.
- Prove that ( \log(z^w) = w \log z + 2\pi i n ), where ( n ) is an integer.
- Prove that ( (z^w)^\alpha = z^{w\alpha} e^{2\pi i n \alpha} ), where ( n ) is an integer.
Strategy: For part (a), we’ll use the definition and the fact that . For part (b), we’ll use the result from part (a) and the definition of complex exponentiation.
Solution:
- Since ( z^w = e^{w \log z} ):
- Compute:
∎
Problem 1.44: Conditions for De Moivre’s Formula
Section titled “Problem 1.44: Conditions for De Moivre’s Formula”Problem.
- If ( \theta ) and ( a ) are real numbers, ( -\pi < \theta \leq +\pi ), prove that
- Show that, in general, the restriction ( -\pi < \theta \leq +\pi ) is necessary in (i) by taking ( \theta = -\pi ), ( a = \tfrac{1}{2} ).
- If ( a ) is an integer, show that the formula in (i) holds without any restriction on ( \theta ). In this case it is known as De Moivre’s theorem.
Strategy: We will use the fact that and the definition of complex exponentiation. For part (b), we’ll provide a specific counterexample. For part (c), we’ll use the fact that integer powers don’t have branch cut issues.
Solution:
- Since ( \cos \theta + i \sin \theta = e^{i\theta} ):
- For ( \theta = -\pi ), ( a = \frac{1}{2} ):
The restriction ensures the principal branch.
- For integer ( a ), ( (e^{i\theta})^a = e^{i a \theta} ), and multiples of ( 2\pi ) cancel, so the formula holds for all ( \theta ).
∎
Problem 1.45: Deriving Trigonometric Identities from De Moivre’s Theorem
Section titled “Problem 1.45: Deriving Trigonometric Identities from De Moivre’s Theorem”Problem. Use De Moivre’s theorem (Exercise 1.44) to derive the trigonometric identities
valid for real ( \theta ). Are these valid when ( \theta ) is complex?
Strategy: We will use De Moivre’s theorem to expand , then equate the real and imaginary parts to obtain the desired identities. Since and are analytic functions, these identities extend to complex .
Solution: By De Moivre’s theorem:
Expand:
Equate parts:
These hold for complex ( \theta ), as ( \cos z ) and ( \sin z ) are analytic.
∎
Problem 1.46: Tangent of Complex Numbers
Section titled “Problem 1.46: Tangent of Complex Numbers”Problem. Define ( \tan z = \frac{\sin z}{\cos z} ), and show that for ( z = x + iy ),
Strategy: We will use the expressions for and in terms of real and imaginary parts, then rationalize the denominator by multiplying by the complex conjugate and simplify using trigonometric and hyperbolic identities.
Solution: For ( z = x + iy ):
Compute:
Multiply by the conjugate of the denominator:
Thus:
∎
Problem 1.47: Solving Cosine Equation
Section titled “Problem 1.47: Solving Cosine Equation”Problem. Let ( w ) be a complex number. If ( w \ne \pm 1 ), show that there exist two values ( z = x + iy ) with ( \cos z = w ) and ( -\pi < x \leq \pi ). Find such ( z ) when ( w = i ) and ( w = 2 ).
Strategy: We will use the expression for in terms of real and imaginary parts, then solve the resulting system of equations for and . We’ll provide specific solutions for the given values of .
Solution: For ( z = x + iy ), ( \cos z = \cos x \cosh y - i \sin x \sinh y = w = u + iv ). Solve:
Square and add:
Since ( w \neq \pm 1 ), solutions exist, with two ( x ) in ( (-\pi, \pi] ).
Case 1: ( w = i ). ( u = 0 ), ( v = 1 ):
For ( x = \frac{\pi}{2} ), ( \sinh y = -1 \implies y = -\ln(1 + \sqrt{2}) ). For ( x = -\frac{\pi}{2} ), ( \sinh y = 1 \implies y = \ln(1 + \sqrt{2}) ). Solutions: ( z_1 = \frac{\pi}{2} - i \ln(1 + \sqrt{2}) ), ( z_2 = -\frac{\pi}{2} + i \ln(1 + \sqrt{2}) ).
Case 2: ( w = 2 ). ( u = 2 ), ( v = 0 ):
Thus, ( x = 0 ), ( \cosh y = 2 \implies y = \pm \ln(2 + \sqrt{3}) ). Solutions: ( z_1 = i \ln(2 + \sqrt{3}) ), ( z_2 = -i \ln(2 + \sqrt{3}) ).
Visualization:
[TikZ diagram omitted in web version]
This visualization shows the solutions to for (red points) and (blue points) within the strip . The gray shaded region represents the strip where we seek solutions.
∎
Problem 1.48: Lagrange’s Identity and the Cauchy–Schwarz Inequality
Section titled “Problem 1.48: Lagrange’s Identity and the Cauchy–Schwarz Inequality”Problem. Prove Lagrange’s identity for complex numbers:
Use this to deduce a Cauchy–Schwarz inequality for complex numbers.
Strategy: We will expand both sides of the identity and show they are equal by careful algebraic manipulation. Since the right-hand side includes a sum of squares of absolute values, it is non-negative, which will immediately give us the Cauchy-Schwarz inequality.
Solution: We want to prove the identity:
It is easier to prove the equivalent formulation:
Let’s expand the left-hand side (LHS):
Now, let’s expand the right-hand side (RHS). The first term is:
The second term on the RHS is:
Adding the two expanded terms of the RHS:
The cross terms cancel perfectly. Comparing the final expressions for the LHS and RHS, we see they are identical. This proves Lagrange’s identity.
To deduce the Cauchy-Schwarz inequality, note that the term is a sum of squares of absolute values, so it must be non-negative. From the original identity, this implies:
∎
Problem 1.49: Polynomial Identity via DeMoivre’s Theorem
Section titled “Problem 1.49: Polynomial Identity via DeMoivre’s Theorem”Problem.
- By equating imaginary parts in DeMoivre’s formula, prove that
- If ( 0 < \theta < \pi/2 ), prove that
where ( P_m ) is a polynomial of degree ( m ) given by
Use this to show that ( P_m ) has zeros at the ( m ) distinct points ( x_k = \cot^2 \left( \frac{\pi k}{2m+1} \right) ) for ( k = 1, 2, \dots, m ).
- Show that the sum of the zeros of ( P_m ) is given by
and that the sum of their squares is given by
Note. These identities can be used to prove that
(See Exercises 8.46 and 8.47.)
Strategy: We will use De Moivre’s theorem to expand and extract the imaginary part. For part (b), we’ll factor out and identify the polynomial. For part (c), we’ll use Vieta’s formulas to relate the coefficients to the sums of roots and their powers.
Solution:
- By De Moivre’s theorem:
Imaginary part:
- For ( n = 2m+1 ):
Zeros at ( \sin((2m+1)\theta) = 0 ), i.e., ( \theta_k = \frac{\pi k}{2m+1} ), so ( x_k = \cot^2 \left( \frac{\pi k}{2m+1} \right) ).
- Sum of roots:
Sum of squares uses trigonometric identities, yielding:
∎
Problem 1.50: Product Formula for ( \sin )
Section titled “Problem 1.50: Product Formula for ( \sin )”Problem. Prove that
for all complex ( z ). Use this to derive the formula
Strategy: We will use the fact that the th roots of unity are the solutions to , then factor out the root and evaluate the resulting product at to obtain the desired formula.
Solution: The roots of ( z^n - 1 = 0 ) are ( e^{2\pi i k/n} ), ( k = 0, \ldots, n-1 ). Excluding ( z = 1 ):
At ( z = 1 ), the left-hand side is ( n ), and:
Thus:
Visualization for n = 6:
[TikZ diagram omitted in web version]
This visualization shows the 6th roots of unity on the unit circle. The blue arrows show the distances from to the other roots, which are related to the sine values in the product formula. For , the product equals .
∎
Solving and Proving Techniques
Section titled “Solving and Proving Techniques”Proving by Contradiction
Section titled “Proving by Contradiction”- Assume the opposite of what you want to prove
- Show this leads to a logical contradiction
- Conclude the original statement must be true
Used in Problems 1.1, 1.9, 1.14, 1.16, 1.17, 1.18, 1.20, 1.22
Mathematical Induction
Section titled “Mathematical Induction”- Verify the base case (usually )
- Assume the statement holds for (inductive hypothesis)
- Prove it holds for using the hypothesis
- Conclude it holds for all positive integers
Used in Problem 1.5 with strong induction
Proving Irrationality
Section titled “Proving Irrationality”- Assume the number is rational (express as )
- Square both sides to eliminate square roots
- Show this leads to a contradiction (usually that a prime divides both numerator and denominator)
Used in Problems 1.9, 1.14
Finding Supremum and Infimum
Section titled “Finding Supremum and Infimum”- For finite sets: find maximum and minimum values
- For intervals: use the endpoints
- For infinite sets: analyze limiting behavior
- For sets defined by inequalities: solve the inequalities to find bounds
Used in Problems 1.19, 1.20, 1.21
Proving Inequalities
Section titled “Proving Inequalities”- Use known inequalities (Triangle, Cauchy-Schwarz, etc.)
- Complete the square or use algebraic manipulation
- Consider cases based on signs of variables
- Use the fact that squares are non-negative
- Used in Problems 1.24, 1.25, 1.26, 1.32, 1.33, 1.48
Working with Complex Numbers
Section titled “Working with Complex Numbers”- Use and powers of cycle every 4
- For division, multiply numerator and denominator by complex conjugate
- Use and (with quadrant adjustments)
- Use De Moivre’s theorem:
- Used in Problems 1.27, 1.28, 1.29, 1.30, 1.31, 1.32, 1.33, 1.34, 1.35, 1.36, 1.37, 1.38, 1.39, 1.40, 1.41, 1.42, 1.43, 1.44, 1.45, 1.46, 1.47, 1.48, 1.49, 1.50
Proving Uniqueness
Section titled “Proving Uniqueness”- Assume two different objects satisfy the same conditions
- Show they must actually be equal
- Often use contradiction or direct comparison
Used in Problems 1.17, 1.18
Using the Pigeonhole Principle
Section titled “Using the Pigeonhole Principle”- Divide a set into fewer subsets than elements
- Show at least one subset must contain multiple elements
- Use this to find numbers that are close together
Used in Problem 1.15
Proving Existence
Section titled “Proving Existence”- Construct an explicit example
- Use intermediate value theorem or similar existence results
- Show that assuming non-existence leads to contradiction
Used in Problems 1.11, 1.15, 1.16
Working with Series and Sums
Section titled “Working with Series and Sums”- Use telescoping series where terms cancel
- Apply geometric series formulas
- Use binomial theorem for expansions
- Factor polynomials and use partial fractions
- Used in Problems 1.2, 1.17, 1.40, 1.49, 1.50
Proving Geometric Properties
Section titled “Proving Geometric Properties”- Use coordinate geometry and distance formulas
- Apply properties of circles, lines, and triangles
- Use complex numbers to represent geometric objects
- Show that conditions imply specific geometric configurations
- Used in Problems 1.13, 1.30, 1.31, 1.34
Using Trigonometric Identities
Section titled “Using Trigonometric Identities”- Apply double angle, sum, and difference formulas
- Use De Moivre’s theorem to derive new identities
- Express complex functions in terms of real and imaginary parts
- Use periodicity and symmetry properties
- Used in Problems 1.44, 1.45, 1.46, 1.47, 1.49, 1.50
Proving Ordering Properties
Section titled “Proving Ordering Properties”- Check trichotomy (exactly one of , , holds)
- Verify transitivity ( and implies )
- Test invariance under operations (addition, multiplication)
- Provide counterexamples when axioms fail
Used in Problems 1.36, 1.37
Working with Logarithms and Exponentials
Section titled “Working with Logarithms and Exponentials”- Use and
- Remember that complex logarithms have multiple branches
- Use for complex exponentials
- Apply properties of complex powers carefully
- Used in Problems 1.38, 1.39, 1.41, 1.42, 1.43
Proving Polynomial Identities
Section titled “Proving Polynomial Identities”- Expand both sides and show they are equal
- Use roots of unity and factorization
- Apply Vieta’s formulas to relate coefficients to roots
- Use polynomial division and remainder theorem
- Used in Problems 1.23, 1.40, 1.49, 1.50