In this course we will study proof writing based on abstract algebra.
Algebra comes from the arabic word “al-jabr" which means reunion of broken parts. In mathematics this word is used for naming a broad area of study which deals with manipulating abstract symbols and objects in a similar manner to that in which we manipulate numbers. Therefore, we shall start by studying the properties of numbers (specifically integers) and later we will build different algebraic structures which obey the same properties.
At the same time we will learn how to write mathematical proofs. This is a skill that takes some amount of patience to master. You might be surprised by the fact that advanced mathematics and in particular proof writing does not consist of a long string of formulas. Instead it consists of many words with some formulas sprinkled throughout. As is the case with a good essay, you might have to go through several attempts and subsequent drafts to write a really good proof.
One very important word of advice about proofs: in this course we will be developing the material from the ground up. This means that when you write a proof you should be able to justify every step relying only on facts previously mentioned in this class and not on any other knowledge.
What is a number? Certainly the things used to count sheep, money, etc. are numbers: .
We will call these the positive whole mumbers natural numbers and write for the set of all natural numbers What you see above is set notation which one reads as “ the set of natural numbers, , consists of 1,2,3,4, etc”.
Since we like to keep track of debts too, we’ll allow negatives and , which gives us the the integers. The word integer refers to a whole number, either negative, positive, or zero. So, the set of all integers is: The symbol is used for the set of integers since the German word for number is Zahlen.
From now on, for brevity, we write to mean is an integer’. The symbol means “in”, as in “belongs to” or “is an element of", so reads “ is an element of the set of all integers”.
Integers have many nice mathematical properties one could ask for: they have operations of addition, subtraction, and multiplication. These operations follow several rules which we will take for granted. We call these facts that we take for granted “axioms".
Axioms 1.1 (Properties of integers).
Closure under addition: If are integers then is an integer.
Associativity of addition: If are integers then .
Commutativity of addition: If are integers then .
Additive identity: there exist an integer so that if is an integer then .
Additive inverses: If is an integer then there exists an integer so that .
Closure under multiplication: If are integers then is an integer.
Associativity of multiplication: If are integers then .
Commutativity of multiplication: If are integers then .
Multiplicative identity: There exists an integer so that if is an integer then .
Distibutivity: If are integers then .
Remark 1.1. If we consider instead the set of natural numbers we see that while most of the above listed axioms continue to hold for the natural numbers, two of them do no. Indeed the additive identity axiom is not satisfied since is not a natural number. Moreover the additive inverses axiom is not satisfied since if is a natural number then is no longer a natural number.
Multiplication produces a relationship which we call divisibility among integers.
Definition 1.2. An integer divides an integer if there exists an integer such that .
Notation 1.1. We write for divides . This should not be confused with the fraction which indicates division.
Example 1.3.
The integers that divide are and — don’t forget the negative ones!
Every integer divides : If is any integer, then since , by definition divides .
The only integer divides is itself: since (for any integer ), by definition divides . But if is any non-zero integer, then for all possible . So, does not divide when .
The only integers that divide are and . However divides any integer since .
The only integers that divide are and . However divides any integer since where is an integer by the additive inverses axiom.
Definition 1.4. An integer is even if divides . In other words, is even provided there exists an integer such that .
Definition 1.5. An integer is odd provided it is not even.
Let us now work out our first mathematical proof!
We will prove the assertion "The sum of two even integers is even." It helps to first rewrite this in if-then form as “ If and are even integers then is an even integer."
The proof below is an example of a proof technique called direct proof, which works best for statements of the form “If P, then Q." A direct proof for such a statement starts with "Suppose P". Then several steps may follow that use axioms, previously proven facts, and formula manipulations. The proof ends when through these deductions you have reached the conclusion that Q is true. This is marked by the proof end symbol whiich looks like a square.
Proposition 1.6. If and are even integers then is an even integer.
Proof. Suppose and are both even integers. By definition of even this means there exists an integer such that and there exists an integer such that and . It follows that Since and are integers, is an integer by closure of the integers under addition. Since we have found an integer, , such that , by definition divides , so is even. ◻
Proposition 1.7. If are integers such that divides and divides , then divides .
Proof. Suppose divides and divides . By definition, this means that there exists an integer so that and there exists an integer such that . We have Since are integers, is also an integer by closure of the integers under addition and mutiplication. Since and is an integer, divides by definition. ◻
Let’s now prove that the sum of an odd integer and an even integer is odd.
We will do so by contradiction. To understand this proof technique consider the following scenario. Suppose I wanted to prove that it is not night outside. If I can look out the window and see the sun then it cannot be night, so it must be daytime. Here seeing the sun contradicts that it is night, so it allows to deduce that it is day.
The proof below is an example of the proof technique called proof by contradiction, which works best for statements of the form “If P, then Q." where is a negative statement. In the proof below will be the statement is odd, which is the same as " is not even". Not is the statement " is even".
A proof by contradiction for such a statement starts with "Suppose P. Assume (towards a contradiction) not Q." Then several steps may follow that use axioms, previously proven facts, and formula manipulations. The proof ends when through these deductions you have reached a contradiction. At this point we say “Since we have reached a contradiction the assumption not must be false, thus is true.
Proposition 1.8. If is an even integer and is an odd integers, then is odd.
Proof. Suppose is an odd integer and is an even integers and assume that is even. Then and by definition of even. By definition of divisibility, there exit integers so that and .
We have Since is an integer, so is its additive inverse . Since and are integers, is an integer by closure under addition. Since and is an integer, divides , so is even by definition. This contradicts that given fact that is an odd integer.
Since we have reached a contradiction, the assumption that is even must be false, thus is odd. ◻
Sets are just collections of objects. the objects in a set are called elements of that set.
Notation 1.2. We write to mean that is an element of a set .
Example 1.9. Writing means is the set with elements 1,2, and 3.
Example 1.10. The empty set, denoted is the set with no elements.
Notation 1.3. We can describe the elements of a set by writing down the properties they satisfy. This is called set-builder notation.
Example 1.11. For example, the set of all even integers is and the set of all odd integers is The vertical line in set builder notation is read "such that". It does not mean divides, although it is the same symbol.
New sets can be obtained from old by means of several set operations.
Definition 1.12. Given two sets and , we can form new sets by
| union | ||
| intersection | ||
| set difference |
Example 1.13. What are ?
Example 1.14. Let and be integers and let .
Can be the empty set? No. Since divides any integer, no matter what and are. So is not the empty set.
Can be a set with only one element? No. Since also divides any integer, has at least two elements ( and ) no matter what and are. So is not a set with only one element.
Can ? Yes, when every integer divides both and .
We can form compound statements as listed below:
and is a true statement whenever both and are true statements
or is a true statement whenever either one of or (or both) are true statements
Not is a true statement whenever is a false statement.
If then (written ) is true whenever is false or is true.
Example 1.15. Which of the following statements are true?
“If pigs could fly then they would be purple.” True since P=“pigs could fly" is false.
“If is even, then divides .” False since P=“12 is even" is true but Q=“ is false.
“For an integer , if is even then divides ”. True. When P=“ is even" is true then Q=“ divides " is true.
The following negation rules are useful. Note that negation takes and to or and vice-versa.
De Morgan’s Laws for negation:
not(P and Q)=(not P) or (not Q)
not(P or Q)=(not P) and (not Q)
We summarize the rules of logic above in truth tables where stands for true and stands for false.
| P | Q | P and Q | P or Q | not P | (not P) Q | (not Q) (not P) | |
|---|---|---|---|---|---|---|---|
| T | T | T | T | F | T | T | T |
| T | F | F | T | F | F | T | F |
| F | T | F | T | T | T | T | T |
| F | F | F | F | T | T | F | T |
Two statements are logically equivalent if they have the same truth value for all possible truth values of the inputs . If statements and are logically equivalent and we want to prove is true, we can prove is true instead!
Remark 1.16. We see from the table that the following pairs of statements are logically equivalent since they have the same truth values (the columns that correspond to the two statements have the same entries):
“If then " is equivalent to “If (not Q) then (not P)".
“ or " is equivalent to “(not ".
Contrapositives and converses
Definition 1.17. The contrapositive of the if-then statement “If then ” is “If not then not ”.
By the second bullet point in 1.16 the contrapositive is logically equivalent to the original statement. Sometimes when proving an if-then statement, it works a bit better to give a proof of the contrapositive. If you can prove the contrapositive is true, it follows automatically that the original statement is also true because of logical equivalence.
Example 1.18. Consider the statement “For an integer , if is odd, then is odd."
If we want to prove this statement we can instead prove the contrapositive “For an integer , if is even, then is even." This latter statement is easier to prove.
Definition 1.19. The converse of the if-then statement “If then ” is “If then ”.
Unlike the contrapositive, the converse of an if-then statement is not logically equivalent to the statement itself.
Example 1.20. Consider the statement: “If Math 310 is in session, the it is daytime." This statement is true. The converse is the statement “If it is daytime, then Math 310 is in session." This statement is false (there is no Math 310 in session at 8am). We see thus that the original statement is not logically equivalent to its converse since they don’t have the same truth value.
In mathematics we use two quantifiers
the universal quantifier “for every” (or “for any" or “for all"), in symbols and
the existential quantifier “there exists”, in symbols .
An important difference
To prove an existentially quantified statement “There exists such that holds." it suffices to give an example of an for which holds.
To disprove a universally quantified statement “For all , holds." it suffices to give a counterexample of an for which is false.
| quantifier | prove | disprove |
|---|---|---|
| proof | counterexample | |
| example | proof |
Example 1.21. Consider the following statement: “The sum of two odd integers is odd." Note that the statement involves implicit universal quantifiers. It implicitly states: “For all odd integers and the sum is odd." According to the table above to disprove this statement it suffices to give a counterexample are two odd integers such that is even.
Negating statements containing quantifiers:
the negation of “For every , holds” is “There exists such that not holds”.
the negation of “There exists such that holds” is “For every , not holds”.
Example 1.22. Negate the following statements:
The sum of two odd integers is odd.
Negation: There exist odd integers and such that is even.
There exists an odd integers so that is even.
Negation: for every odd integer , is odd.
First we look at a special property of the non-negative integers.
Axiom 1.23 (The Well-Ordering Principle). 1 Every non-empty set of non-negative integers has a smallest element.
Understanding the Well-Ordering Principle.
Example 1.24. For example, the set is a non-empty set of non-negative integers, and so the Well-Ordering Principle tells us that it has a smallest element. In this example, it’s evident: .
This also holds for infinite sets: Let be the set of all strictly positive multiples of . It has a smallest elements (namely, ). The principle also applies even when it is not obvious what the smallest element is. For instance, consider the height measured in inches (and rounded down to the nearest whole number) of all giraffes on planet Earth. That is, let This set must have a smallest number (there is a shortest giraffe — poor guy!).
Remark 1.25. The word “non-empty” is crucial. The statement “Every set of non-negative integers has a smallest element” is false, since the empty-set (which is considered a subset of the set of non-negative integers — the empty set is a subset of every set) has no smallest element, since it doesn’t have any elements at all.
Remark 1.26. The word “non-negative” is also important. For instance, the set of all negative multiples of , namely is a non-empty set of integers, but it has no smallest element.
Let us now discuss the so-called “division algorithm”:
Theorem 1.27 (The Division Algorithm Theorem). Let with . There exists unique such that
Example 1.28. If we take and in the statement of the Division Algorithm, then the values of and that make the conclusion hold true are and : and .
Let us check that and are the only values that work: Suppose and . So is one of , or . If , then , but does not divide , and so is not possible. If , then and thus . But does not divide , and so is not possible. If , then and thus . But does not divide , and so is not possible. We thus must have that . Then and hence , so that must be .
Example 1.29. If we take and in the statement of the Division Algorithm, then the values of and that make the conclusion hold true are and : and .
Remark 1.30. If we were to allow to be negative, the conclusion of the Division Algorithm Theorem would fail to be true: note that it is impossible for any integer to satisfy when , so no such integers and would exist in this case.
Remark 1.31. If we were to drop the phrase “" from the Division Algorithm Theorem, it would also become false, since the uniqueness portion would fail: For instance, if and , then but also , so without the clause “", there would not be a unique pair of integers and that satisfy the requirements.
The Division Algorithm Theorem has two parts: an existence statement and an uniqueness statement. Existence says that there is at least one way of writing with . The uniqueness part says there is at most one way of doing so. In other words, the uniqueness clause asserts that if we have with and we also have with , then it must be that and .
Our next goal is to prove the existence portion of the division algorithm. Let’s recall what this says exactly:
Theorem 1.32 (Existence portion of the Division Algorithm).
Let be integers with . There exist integers and such that and .
Before giving the formal proof, let’s think intuitively for a bit. If I want to divide by , for instance, leaving a remainder in the range , we could proceed by subtracting off ’s from repeatedly until it is no longer possible to do so without going into negative territory. That is, the remainder is what’s left over after I subtract off as many ’s as a I can from — in this example the remainder is .
In general, given and , we want to subtract from as many copies of as possible. (Things are more complicated when is negative, but don’t worry about this for now.) This motivates the definition of the set So, an element belongs to if (1) a positive integer and (2) is the result of subtracting from some number of copies of .
For instance if and then . Note that in this example the smallest element of is , which is indeed the remainder of division of by . We can thus “find” the remainder by seeking out the smallest element of , and this turns out to work in general. Moreover, once we know what the remainder is, the value of is determined by the equation .
Let us now give a formal proof:
Proof of Theorem 1.32. Let and be integers with . Define a set of non-negative integers as follows: By construction, consists of only non-negative integers. We would like to apply the Well-Ordering Principle, but to do so we must check that is not empty.
Let us prove is non-empty by considering two cases. If , then taking in the formula above gives , which belongs to (since it is non-negative by assumption). If , then taking in the formula above gives . Since , we have , and since , it follows that . Thus is an element of . In either case, we have shown is not the empty set.
Since is a non-empty set of non-negative integers, by the Well-Ordering Principle, has a smallest element, which we will call . (Our aim is to prove that this is the that “works” to give the conclusion.)
By definition of , we have that and that for some integer . Let us now show that also. Suppose this is not the case; that is, suppose . Since , it follows that . Since , we have . This shows that . But , which contradicts our assumption that is the smallest element of . We conclude that it must be the case that .
We have shown that there are integers and such that and . Rewriting the first equation slightly, we have shown that there are integers and such that and . ◻
Let us prove the “uniqueness” portion of the Division Algorithm. Here is what it says:
Theorem 1.33. Let be integers with . Suppose that with and with . Then and .
Proof. We first show that divides : Since and , we have and hence . Since is an integer, this shows that divides .
We next show that : Since , we have , and since , wehave . So . Similarly, since we have and since , it follows that , whence . Putting these together gives
Now, the only integer such that and divides is . Since we have shown that has these two properties, it follows that ; that is, .
Finally, since and , we conclude that . Since , it follows that and thus . ◻
Definition 1.34. Let be an integer. A divisor of is an integer so that divides .
Definition 1.35. Let and be two integers, not both of which are . The greatest common divisor or GCD of is the largest integer such that divides and divides . We often write for the GCD of and .
More formally by definition has the following properties:
divides both and . (“common divisor" property)
If is an integer that divides both and , . (“greatest" property)
According to the above definition if you want to prove that an integer is the GCD of and you must show
divides both and . (“common divisor" property)
If is an integer that divides both and , . (“greatest" property)
Example 1.36. For example, let us find by listing all the common factors of and and finding the largest one: The factors of are . The factors of are So the common factors are . The largest of these numbers is ; whence .
Example 1.37.
For , what is ? Since divides itself and nothing bigger than does, . For , what is ? When , is the largest divisor of so .
What is when ? Since divides both and and nothing bigger than divides , we have . What is when ? Since divides both and and nothing bigger than divides , we have .
What is for ? Every integer divides . So is the largest divisor of . For , the largest divisor of is , but if , then the largest divisor of is . So .
Lemma 1.38. If with , then .
Proof. Suppose with . We first show that divides if and only if divides . If then there exists so that and thus shows that . Conversely if then there exists so that and thus shows that .
So, the set of common divisors of and is identical to the set of common divisors of and . By definitions, is the largest element of the first set and is the largest element of the second set. Since these sets are identical, . ◻
Theorem 1.39. Let and be integers with , and suppose are integers which satisfy . Then .
Proof. Suppose and are integers with . First, suppose that is an integer that divides both and . Then we can find integers and such that and , and Since are integers by closure of the integers under addition and multiplication . Therefore, divides , and by assumption also divides . Now since is the largest integer that divides both and , we must have by property (2) in Definition 1.35. In particular, this applies to the special case when , since that is an integer that divides both and . We conclude that .
Now we will show that . To do that, suppose now that is an integer that divides both and . This means we can find integers and such that and , so Again since are integers by closure of the integers under addition and multiplication . This shows that must also divide . Since by assumption also divides , we must have by propery (2) in Definition 1.35. This applies in particular to the case when , so we conclude that .
We have shown that and , and hence . ◻
Theorem 1.39 is used in practice to find, in an efficient manner, the gcd of two integers. Let’s look at an example:
Example 1.40 (Euclidean algorithm). Suppose we want to find the gcd of and . Consider the following calculations: In each step, we are applying the Division Algorithm, starting by dividing by to leave a remainder of , then dividing by , leaving a remainder of , and so on. In general, if in one step we are dividing by leaving a remainder of , in the next step we are dividing by , leaving a new remainder.
By Theorem 1.39, if , then . Applying Theorem 1.39 to equation [1] it shows that . Applying Theorem 1.39 to equation [2] it shows that , and continuing to apply Theorem 1.39 to the remaining equations we obtain Since by Example 1.37, the equalities above give .
The Extended Euclidean Algorithm
In addition to finding the gcd of two integers and , the Euclidean algorithm can also be used to find integers and such that . This is sometimes called the “Extended Euclidean Algorithm”.
Definition 1.41. An expression of the form is known as a integer linear combination of and .
With this terminology, the extended Euclidean algorithm gives a method of expressing as an integer linear combination of and . Let’s see how this works in the example above:
Example 1.42 (Extended Euclidean algorithm). Our next goal is to express as a linear combination of and . We can do so re-using the work from the Euclidean Algorithm as follows:
Here are formal statements of the Euclidean Algorithm and its extended version.
Theorem 1.43 (The Euclidean Algorithm). Let and be integers such that . Define a list of numbers such that as follows:
define to be the unique integer in the Division Algorithm such that for some integer and . (If , then the process stops at .)
If , define to be the unique integer in the Division Algorithm such that for some integer and . (If , then the process stops at .)
If , define to be the unique integer in the Division Algorithm such that for some integer and . (If , then the process stops and .)
If , define to be the unique integer in the Division Algorithm such that for some integer and . (If , then the process stops and .)
Continue in this fashion until one arrives at an integer that is . More precisely, .
Then .
Proof. Claim 1: the algorithm terminates in finitely many steps.
The inequalities from the Division Theorem can be combines as follows Since there are finitely many integers between and which can be the successive remainders in the algorithm, the process described in the Theorem cannot continue forever; that is, for some . In fact, notice that the process must stop after at most steps.
Next we want to show
Claim 2: .
Since , by Theorem 1.39 we have . Moreover, since , Theorem 1.39 also implies that . Since , by Theorem 1.39 we must also have . Continuing in the manner, applying Theorem 1.39 for each step of the algorithm, we obtain the list of equations Recall that , and so from the equality of the first and last quantities above we have . Since for any non-zero integer , we have . This proves that . ◻
A consequence of the Extended Euclidean Algorithm is
Theorem 1.44. If are integers, , and , then there exist integers so that .
Definition 1.45. A integer is called prime if
, , and
the only divisors of are , , and .
Definition 1.46. An integer is composite if it is not prime.
Lemma 1.47. If is prime and is any integer, then is either or . In particular, if does not divide , then .
Proof. Suppose is prime and is any integer. By definition, the only divisors of are , , , and . If divides , then so does , and since is the largest divisor of , we conclude that . If does not divide , neither does , so is the only positive integer dividing both and , and . ◻
Theorem 1.48. Let be integers. If divides and , then .
Proof. Suppose are integers, divides and . By the extended Euclidear Algorithm (Theorem 1.43) the greatest common divisor of and is always an integer linear combination of and ; specifically there exist integers and such that . Therefore, multiplying by gives On the other hand, our assumption that divides says that there exists an integer such that . Therefore, Since is an integer by closure of under addition and multiplication, by definition, this means that divides . ◻
Corollary 1.49. Assume are integers. If is a prime and divides , then must divide or .
Note that “or" here is not exclusive: the theorem says that one of the following three statements is true:
either divides a,
divides ,
or divides both and .
Proof. Assume is prime and divides . We consider two cases. If divides , we are done since we already have what we wanted to prove. If does not divide . By Lemma 1.47, . Since and we assumed that divides , we can apply Theorem 1.48, which leads us to conclude that must divide . ◻
Corollary 1.50. Suppose is a prime integer, are arbitrary integers, and divides . Then divides at least one of or .
Proof. The most rigorous way to prove this would be a proof by induction, a topic that has not yet been discussed in this class. We will make do with a slightly less formal proof:
Assume is prime and divides . Thinking of the latter product as , by Corollary 1.49 we have that divides or divides . If divides , then we are done – we have found an that divides. If not, then divides . Now applying the Corollary again gives that divides or divides . Likewise, if divides , then divides or divides . Continuing in this manner, we arrive at the fact that divides or divides or or divides . ◻
Corollary 1.51. Suppose , are all prime integers, and that divides . Then for some .
Proof. By Corollary 1.50, we have that divides for some . Since is prime, its only divisors are and . But and (by the definition of “prime”) and hence it must be that . ◻
The Fundamental theorem of arithmetic says that every integer can be factored into primes in an essentially unique way. More precisely, it says the following:
Theorem 1.52 (The Fundamental Theorem of Arithmetic). Any integer can be written as a product of primes. Moreover, if are two factorizations of into primes, then , and there exists a reordering of the ’s such that for all .
Example 1.53. For example, we can factor into a product of primes as follows and also as These are “essentially the same” in the sense that if we rearrange the factors and change a couple of the signs, we can tranform one factorization into the other one.
Example 1.54. For another example, is already factored into a product of primes: it is a product of one prime.
The Fundamental Theorem consists of an “existence" and “uniqueness" statement. The existence part says that every nonzero integer can be written as a product of prime integers. More formally, for any integer such that , there is a non-negative integer and a list of primes such that . The uniqueness part says that this list of primes is unique up to signs and the order of the primes. More formally, that given wo factorizations of into prime say then , and we can reorder of the ’s such that for all .
The existence portion of the Fundamental Theorem of Arithmetic states the following:
Theorem 1.55 (FTA existence). For any integer , other than , or , there is a list of prime integers for such that .
Proof of the existence portion of the Fundamental Theorem. Suppose is an integer other than , or .
Case 1: Let us first consider the case when . Let be the set of all integers that are greater than or equal to that are not products of primes: Our goal is to prove is the empty set.
By way of contradiction, assume is not empty. Then, by the Well-Ordering Principle, will have a least element, call it . In other words, is an integer such that and such that cannot be written as a product of primes, but such that every integer such that can be written as a product of primes.
We note that cannot be prime, since a prime integer is its own prime factorization. (Note that is allowed in the statement of the theorem.) So, we must have that for some and . But then since and are both less than and is the least element of , both and must not in . Since each is greater than or equal to , it follows that each of them is a product of primes. Say and where each and is prime. Then is a product of primes, contrary to the fact that . We conclude that must be the empty set — that is, every integer that is at least is a product of prime intgers.
Case 2: Now assume . Then and hence, by the case already established, we have that for some primes . It follows that . Since is prime, so is . This proves is a product of primes. ◻
The uniqueness portion of the Fundamental Theorem states:
Theorem 1.56 (FTA uniqueness). If for some primes and , then and, after possibly reordering the ’s, we have that for all .
Proof. Suppose is an integer other than , or .
Case 1: Let us first consider the case when .
We give a proof by contradiction. Consider the set
are two prime factorizations which differ not just by reordering and negating the factors .
Assume towards a contradiction that is not empty. Since consists of nonnegative integers, by the Well-Ordering Principle, there is a smallest element of — let us call it . Then has two factorizations, say for some primes and , such that no possible reordering and changing of signs can make them be the same.
We can rewrite the equations above as which shows that divides . Since is prime as are , it must be that for some by Corollary 1.51. By renumbering the ’s, we may assume , so that . Dividing by we get with still a positive integer. Since is a prime , and so we must have . Thus and hence the two prime factorizations of above must be essentially the same, that is, and after reordering for . Together with this shows that the two factorizations of are essentially the same, a contradiction.
We conclude that is empty, so every integer has an essentially unique prime factorization.
Case 1: Now assume . Then and hence, by the case already established, since we have that and after possibly reordering we have for . ◻
Lemma 1.57. If the Fundamental Theorem of Arithmetic is true and is an integer then there exist positive prime integers such that .
Proof. Suppose is an integer and the FTA is true.
By the FTA, there exist primes so that . Then we have Since the absolute value of a prime is a prime, each of the integers are positive primes, so by setting we obtain that and are positive primes. ◻
Let’s see a sample application for the uniqueness statement of the FTA .
Example 1.58. Are there primes and so that and ?
No this is not possible. The FTA says that if we have two prime factorizations and that are equal then the number of prime factors in each is the same. This would mean that , a contradiction.
Here is the main result that relates prime factorizations and the notion of divisor.
Lemma 1.59. Let and pe positive integers so that with distinct primes and integers. Then if and only if for some integers so that for all .
With this interpretation of the divisors in terms of prime factorization, we can determine a formula for the GCD.
Theorem 1.60. Let and be positive integers so that with distinct primes and integers. Then where denotes the smallest among and .
Proof. By Lemma 1.59, if and only if for some integers so that and if and only if for some integers so that . Thus is a common divisor of and if and only if and . The largest number of this form is . This number is therefore the largest common divisor . ◻
Example 1.61. Suppose and then we can first write the numbers using all primes that appear in either, using exponents equal to zero if necessary: and . Then the formula in the theorem above gives
Induction is a very helpful proof technique for showing results about all positive integers, or where the statement depends on a positive integer. Usually, proofs by induction go as follows:
We have a statement we want to prove that depends on a positive integer ; let’s refer to that statement as . To prove for all , we follow two steps:
Base Case: Prove that holds for the smallest value of we are considering. Usually this is or .
Induction Step: Assume that is true for some , and prove that this implies is true.
The Domino Metaphor: Imagine that the statements are like dominoes. Suppose I tell you that (a) the dominoes have been arranged so that if one of the them falls over, the next one in line will fall too, and (b) I have knocked over the first one. Then it is clear that, eventually, every domino will fall. (Ignore the fact that it would take an infinite amount of time for this to happen in reality.) The Principle of Mathematical Induction is an abstraction of this idea.
It is a theorem that once you prove both the base case and the induction step, must hold for all integers , where is the value of in your base case. Most often .
Here is the formal statement:
Theorem 2.1 (The Principle of Mathematical Induction). Suppose is an integer and is a statement that applies to an integer . If is true and implies for all , then is true for all .
Proof. Suppose is true and implies for all .
We give a proof by contradiction. Assume it is not the case that is true for all . Then the set is a non-empty set of positive integers. By the Well-Ordering Axiom, has a least element, call it . In words, is an integer such that is true but is false for all such that .
Since we assumed is true, . So and hence . Since , must not be in the set . That is, is true. But we are assuming implies , and thus must be true. We have concluded that ) is both true and false, which is a contradiction.
Since we reached a contradiction, it must be that the set is empty — that is, is true for all positive integers . ◻
Here is a statement that can be proven by the Principle of Mathematical Induction:
Recall that .
Theorem 2.2. For every integer , we have .
Proof. We give a proof by induction. Let be the statement "".
Base case: is the statement . This is true since .
Induction Step: Assume is true for some , that is, .
is the statement
Observe that where the middle inequality follows by multiplying the inequalities (the inductive hypothesis) and (which follows from ).
This shows is true.
By the Principle of Mathematical Induction, for every integer .
◻
When writing a proof by induction, you should tell the reader that you’re about to do a proof by induction, and you should clearly indicate where your base case and induction step are. The assumption that holds for some used in the induction step is called the induction hypothesis. The fact that proving your base case and induction steps is sufficient to prove the theorem is recalled by invoking the Principle of Mathematical Induction to finish the proof.
Important features of a proof by induction:
always start by stating you will do a proof by induction
next state clearly what your statement is
a frequent mistake is to include the words "for all/every/any " in the statement. This is incorrect. The statement applies just to the which is named inside the notation.
label your base case and induction step clearly
state what is clearly before you prove it.
finish the proof saying “By the Principle of Mathematical Induction is true for all .
Here are some more proofs using the Principle of Mathematical Induction:
Theorem 2.3. For all , .
Proof. Let’s prove this by induction on . Let be the statement .
Base Case: When , we do indeed have , so holds.
Induction Step: Suppose that holds for some , so that . Then so holds.
By the Principle of Mathematical Induction, for all . ◻
Example 2.4. For every integer , divides .
Proof. We give a proof by induction. Let be the statement "5 divides ".
Base case: for , is the statement that divides .
Induction Step: Assume is true for some . This means that 5 divides , so by definition of divisibility there exists so that or .
We wish to prove that is true. is the statement that divides . Observe that Since , then by closure under addition and multiplication. Since and , divides . So is true.
By the Principle of Mathematical Induction, 5 divides for all integers .
◻
We can also reprove the following theorem using induction on :
Theorem 2.5. Let be integers. If is a prime and divides , then divides for some .
Of course we have already proved this before! But now we can give a more formal proof — our previous proof was a bit handwave-y. We will need the following result, which we have proved before:
Theorem 2.6 (Euclid’s Lemma). If is a prime and divides , then divides or divides .
Proof of Theorem 2.5. Let’s fix a prime . Let be the statement “If divides for some integers , then divides for some ”. We prove holds for all by induction.
Base Case: When , if divides , then divides the only one of the , which is .
Induction Step: Suppose that holds for some , meaning that if divides , then must divide some of the . Now suppose that are integers and divides . Since , we must have , so that we have at least two different . Since divides , Theorem 2.6 says that divides or divides . If divides , we are done. Otherwise, must divide , so by induction hypothesis, divides some for . Either way, divides one of the numbers in the list , and we have shown that holds.
◻
One thing that might be confusing about this proof is that Theorem 2.6 is precisely . We needed this statement in our induction step.
Let’s now prove the uniqueness part of the FTA.
Theorem 2.7 (FTA uniqueness). If for some primes and , then and, after possibly reordering the ’s, we have that .
Proof. We give a proof by induction on . Let be the statement “If and for some primes and , then and, after possibly reordering the ’s, we have that ."
Base Case: Suppose and for some primes and . Since and are positive and we must have and . So and the equality is just . This shows is true.
Induction Step: Suppose is true. Suppose and for some primes and . Since and , we have that . By Theorem 2.5 we conclude that for some . After reordering the the ’s, we may assume that and since both and are primes this gives .
Substituting into we have and dividing by yields Observe that so that applies to the above displayed equality and gives and . Since we have and since from above we have . Together with this shows that is true.
By the principle of mathematical induction, we have proven that the theorem is true. ◻
Sometimes we need a stronger form of induction where in the induction step we assume not only , but also for all , where should also satisfy or whatever else is our base case. This is called Strong Induction or Complete Induction. Here is the formal statement:
Theorem 2.8 (The Principle of Strong Mathematical Induction). Suppose is a statement that refers to an integer . Suppose that there exist integers so that
are true and
for all integers , if is true for all , then is true.
Then is true for all .
There are two differences between strong induction and usual induction. First, in strong induction one typically needs to prove more than one base case (). You will have to include as a base case all the cases for which you cannot prove the induction step. Second, in the induction step of strong induction we get to assume any previous statements with and use whichever one is convenient. This flexibility is what makes this version of induction “strong".
Example 2.9. Use strong induction to prove that every class with at least students in it can be divided up into groups each of size or .
Proof. Let be the statement “A class with students can be divided up into groups each of size of ”. We prove is true for all by strong induction.
Base case: For the base cases, we start by checking each of the cases , , and hold. This is true since , and .
Induction step For the induction step, assume and that is true for all such that . Our goal is to prove is true, that is, a class having students can be divided up into groups each of size of .
From a class of students, pick any three students and put them together into a group. Note that since we have assumed . Also we have . So the strong induction hypothesis applied for gives that is true— that is, the remaining students can be put into groups of size or . In all, we have put all students into groups of size or combining the group of 3 we formed initially with the grups formed form the other . This shows that is true.
By the Principle of Strong Mathematical Induction, is true for all . ◻
Let’s use strong induction to reprove the existence part of the Fundamental Theorem of Arithmetic.
Theorem 2.10. Every integer is a product of primes.
Proof. We give a proof by strong induction. Let be the statement the integer is a product of primes.
Base Case: When , is prime, so it is a product of primes — in fact, it is a product of one prime.
Induction Step: Suppose that every integer is a product of primes. We want to show that is a product of primes.
If is prime, then it is a product of primes, and we are done. If is not prime, by Lemma [lemma composite] we can find integers and such that . In fact, setting and and using that we see that with and and and . This means that and so and .
Since and , by the induction hypothesis we can find primes and such that so is a product of primes.
By the principle of strong mathematical induction every integer is a product of primes. ◻
Definition 3.1. An equivalence relation on a set is a binary relation satisfying the following properties:
For every , . This is known as the reflexive property.
For all , if , then . This is known as the symmetric property.
For all , if and , then . This is known as the transitive property.
Typically the symbol “” is read as “is equivalent to”; is read as “ is equivalent to ”.
Example 3.2. Say is the set of students in our class and a relation is defined on this set as follows: students and are equivalent, written , if they are born in the same month. We check this is an equivalence relation:
For every , is born in the same month as , thuse . Thus the relation staisfies the reflexive property.
If , then and are born in the same month, so and are born in the same month, which shows . This means that the relation satisfies the symmetric property.
If and , then and are born in the same month and and are born in the same month so in fact all there () are born in the same month. In particular, and are born in the same month so . This proves the transitive property.
Since it satisfies the reflexive, symmetric and transitive properties the relation of being born in the same month is an equivalence relation.
Here is an example of a binary relation that is not an equivalence relation:
Example 3.3. Let and set to mean . This relation is symmetric and transitive but not reflexive. We show it is not reflexive by means of a counterexample: we have so , but since .
Here is another example:
Example 3.4. Let denote the set of all points in the plane. Given two such points and , declare to mean that and have the same distance to the origin. For instance the point has distance to the origin and the point also has distance to the origin, and hence . Put and are not similar for .
The three properties of an equivalence relation are pretty clear in this case: (1) Any point has the same distance to the origin as itself, (2) if and have the same distance to the origin, so do and , and (3) if and have the same distance to the origin and and have the same distance to the origin, then and are of the same distance to the origin.
Definition 3.5. Let be an equivalence relation on some set . For , the equivalence class of , written , is the subset of consisting of all elements that are equivalent to :
Example 3.6. Let and be as in Example 3.4. Given a point , what is the equivalence class ? It is the set of all points in the plane located on the circle centered at the origin which goes through .
Suppose the distance from to the origin is while the distance from to the origin is . As above is the circle centered at the origin of radius and is the circle centered at the origin of radius .
If then . Thus the circle centered at the origin of radius is the same as the circle centered at the origin of radius , that is .
If then . Thus the circle centered at the origin of radius and the circle centered at the origin of radius do not share any points, that is .
Theorem 3.7. Let be any set and be any equivalence relation on . For elements we have if and only if .
Proof. If then .
If then since (by the reflexive property) we have too. By definition of , this means .
If then . Suppose . Pick any . By definition this means that . Since , by transitivity, this gives that and hence . We have proven . Now pick any . By definition this means that . Since , by symmetry we also have and hence by transitivity we have , which means that . This proves that . Since and , we have . ◻
Corollary 3.8. Let be and set and an equivalence relation on . Any two equivalence classes are either equal or disjoint: for all , either or .
Since “either or " is a statement of the for “ P or Q" we will instead prove the logically equivalent statement “ not(Q) P" , that is, “if then ".
Proof. Suppose . Then there exists meaning that and . Since we have by definition of equivalence class and since we have by definition of equivalence class. By the symmetric and transitive properties, since and , then . Since , by the previous Theorem we deduce that . ◻
The corollary says that an equivalence relation on a set “partitions” into equivalence classes: Every element of belongs to one, and only one, equivalence class. Equivalently, is the disjoint union of its equivalence classes.
Definition 3.9. Fix a nonzero integer . We say that are congruent modulo if divides . We write for “ is congruent to modulo ". In this notation, the and are the two inputs, and is one piece, like a complicated equal sign.
Here are some examples:
because divides .
is not true, because does not divide .
because divides .
Theorem 3.10. Any two odd integers are congruent modulo 2.
Proof. Let and be two odd integers. By a problem in problem set 1, and for some integers and . Now so divides , which means that . ◻
Example 3.11. It is not true that any two odd integers are congruent modulo 3, because for example (since does not divide ) and and are both odd.
Example 3.12. If for some integer , what are the possible values of ? The congruence holds if and only if divides , and so must be one of the divisors of : .
Congruence classes modulo The number is congruent to modulo , because is divisible by . One way to find this is to consider the division algorithm: dividing 2001 by 10 we obtain which shows that is divisible by 10.
A different way to think about this: for any positive number , if is the digit appearing in it one’s place, then the ones place of is and hence is divisible by . Thus, where is the digit in the one’s place of .
However, we have to be a little more careful about negative integers. If is an integer, to find the unique such that , we consider the last digit of , and take . For example, and .
Theorem 3.13. For a fixed , every is congruent modulo to some such that .
Proof. By the Division Algorithm we have So and thus divides . It follows from the definition of congruence that . ◻
Theorem 3.14. Congruence modulo is an equivalence relation; that is, the following three properties hold for any nonzero integer :
Symmetric: For any integer , .
Reflexive: For all integers and , if , then .
Transitive: For all integers , , and , if and , then .
Proof. Fix a nonzero integer .
For any integer , is divisible by and hence .
For integers and , assume . Then divides and hence also divides . This proves that .
Let , , and be integers such that and . By definition this means divides and divides . Using the fact proven before that if divides two integers then it divides their sum, we see that divides . By definition, this proves .
◻
Theorem 3.15 (Congruences can be added). Given a nonzero integer and integers , if and , then .
Proof. Fix a nonzero integer and assume are integers such that and . Then divides both and and hence also divides their sum, which is . By definition this means that . ◻
Theorem 3.16 (Congruences can be multiplied). Given a nonzero integer and integers , if and , then .
Proof. Fix a nonzero integer and assume are integers such that and . Then divides both and , so and for some .
Therefore where is an integer by closure of under addition. This means that which shows by the definition of congruence.
Since and we have It follows that , where is an integer by closure of the integers under addition and multiplication. Thus and hence by the definition of congruence. ◻
Definition 3.17. Fix a nonzero integer . For , the congruence class of modulo is the subset of consisting of all integers congruent to modulo ; that is, the congruence class of modulo is the set
Let me emphasize that congruence classes are sets, not numbers.
Example 3.18. What are some of the elements in ? Here are some examples and .
The set consists of all the odd integers. The set consists of all the even integers.
True or False? Justify.
. This is true since .
. This is true: an element belonging to both sets would have to be congruent to both and modulo . If such a number existed, then and would be congruent modulo (using the symmetric and transitive properties of congruences). But they are not. So no such number exists.
. This is false; for instance, belongs to both these sets. Notice that we are using two different values for in this example.
For all integers , . This is true. If then and hence divides . But then would also divide and hence , so that .
Example 3.19. Set . Then there are three congruence classes modulo :
Notice that every integer belongs to one, and only one, of these three sets. They partition the integers into three non-overlapping subsets, likes countries on a map. A bit more formally, this property looks as follows.
Theorem 3.20. For any positive integer and any pair of integers exactly one of the following possibilities is true:
or
Proof. Since we have seen in Theorem 3.14 that congruence modulo is an equivalence relation, and since the congruence classes are just the equivalence classes of this equivalence relation, the claim of this theorem follows from Corollary 3.8. ◻
We can make more precise when each of the two cases in the preceding theorem occurs.
Theorem 3.21. For any positive integer and any pair of integers we have
if and only if
if and only if .
Proof. The statement “ if and only if " is true by Theorem 3.7 applied to the equivalence relation congruence modulo .
We will now prove the statement “ if and only if ".
Suppose . Then since we obtain that . Therefore, by definition of congruence classes .
Suppose then but . This shows that , By Theroem 3.21 this means that must be true. ◻
Theorem 3.22. For any positive integer , there are exactly congruence classes modulo , namely the congruence classes .
Proof. Suppose is a positive integer. By Theorem 3.13 we have that any integer is congruent to some integer so that . By Theorem 3.21, since we have for some integer so that . Since , the possibilities for are and thus the possibilities for are . We thus have at most congruence classes modulo .
To see that the congruence classes are distinct, we appeal to the uniqueness of in Theorem 3.13. Indeed, if we hve that so that and then applying Theorem 3.13 for we get that both and satisfy the properties required for and thus by uniqueness. Thus the congruence classes are distinct and so we have exactly congruence classes. ◻
We come to a very important point. We will see shortly that although congruence classes modulo are sets, not numbers, they behave like numbers in the sense that they can be added and multiplied. Thus we can talk about arithmetic on the following set.
Definition 3.23. For a positive integer , is the set of congruence classes of integers modulo , that is
Using the fact that congruences can be added and multiplied, we may introduce rule for adding and multiplying congruence classes
Definition 3.24. Fix a non-zero integer and let . Define to be the congruence class modulo given by the following rule:
Pick any and pick any , and define .
For example if we picked and we would get
Example 3.25. Let us compute . The rule says that we start by picking any elements and . I’ll choose and . Next we add our choices for and as ordinary integers to get , and conclude that . Note that , so we could also say that . Since and , we could rewrite this yet again as .
Now you try it:
Example 3.26. Have each member of your group compute . Then compare your answers. Are they equal? Repeat this for .
Solution: If one of us were to pick and , the rule for adding congruence classes gives . If another one of us were to pick and , the rule for adding congruence classes gives . This gives the same result since , because .
Perhaps you notice a troubling aspect of the definition for adding congruence classes: the definition of seems to depend on choices of representative elements and . What if one pair of choices leads to a different outcome than another pair of choice? That would be non-sensical.
Theorem 3.27. The definition of is independent of the choices mades. That is, if and , then .
Proof. Our proof will use two theorems proven before: (1) and belong to the same congruence class modulo if and only if (Theorem 3.21) and (2) congruence classes may be added (Theorem 3.15).
Since and belong to the same congruence class, by Theorem 3.21 we have that , and likewise since and belong to the same congruence class, . By Theorem 3.15, we may add these two congruence classes to conclude . Using Theorem 3.21 again, it follows that . ◻
Now we can define multiplication of congruence classes.
Definition 3.28. Fix a non-zero integer and let . Define to be the congruence class modulo given by the following rule:
Pick any and pick any , and define .
For example if we picked and we would get
Theorem 3.29. The definition of multiplying congruence classes is independent of the choices made. In more detail, if and are two congruence classes modulo , and we pick , and , then .
Proof. Our proof will use two theorems proven before: (1) and belong to the same congruence class modulo if and only if (Theorem 3.21) , and (2) congruence classes may be multiplied (Theorem 3.16).
Since and belong to the same congruence class, by (1) we have that , and likewise since and belong to the same congruence class, . By (2), we may add these two congruence classes to conclude . Using (1) again, it follows that . ◻
Definition 3.30. A congruence class in is called a unit if it has a multiplicative inverse. A multiplicative inverse of is a congruence class in such that .
Example 3.31. In , and are the only units. The multiplicative inverse of is , and the multiplicative inverse of is .
Theorem 3.32. Given a non-zero integer and any integer , there exists some integer such that if and only if .
Proof. Recall that if and only if for some integers and . The existence of integers and such that is equivalent to the existence of an integer such that divides , or equivalently, such that . ◻
We can translate Theorem 3.32 into a statement about congruence classes:
Corollary 3.33. Given a non-zero integer and any integer , the congruence class is a unit in if and only if .
Proof. There exists some in such that if and only if there is an integer such that if and only if by the previous theorem. So the congruence class is a unit if and only if . ◻
How do we find multiplicative inverses? We can use trial and error, but it gets very slow; or we can use the extended Euclidean algorithm.
Example 3.34. Note that is a unit. We can find its multiplicative inverse via the extended Euclidean algorithm, as follows. First, we note that so that indeed . We can now use the computations above to find a linear combination of and that is equal to 1:
Therefore, so that is the multiplicative inverse of .
Definition 4.1. A binary operation on a set is a function from to . In other words, it is a rule that assigns to two inputs taken from , one output which also belongs to .
Example 4.2. For example, addition and subtraction are binary operations on set the of integers.
If denotes a binary operation, we write to indicate the result of applying to the pair of inputs , just as we would with the usual symbols and for the sum and subtraction of integers.
Remark 4.3. In defining a binary operation, we are implicitly stating that the set is closed under the operation.
Example 4.4. For instance, if we tried to define on the set of non-zero integers, we would run into trouble: this is not an operation since isn’t always an integer.
Example 4.5.
Which of the following rules are binary operations on the indicated set ?
Let be all non-zero real numbers and let .
Let be the set of all even integers and
Let be the set of all odd integers and
Solution: The operations in (a) and (b) are in fact binary operations in the sets given, though (c) is not since it does not satisfy closure: the sum of two odd integers is never an odd integer.
Here are some special abstract properties a binary operation can have.
Definition 4.6.
Commutativity. A binary operation is commutative if for any .
Associativity. By definition, binary operations only take two inputs. If we wanted to operate on three things, we would have to choose two to pair first, then throw in the third. A binary operation is associative if we get the same result with either grouping:
Identity. An element is an identity for a binary operation if and for all .
Inverses. If is a binary operation with an identity , then an inverse for an element is another element such that and .
Example 4.7. For each of the following binary operations, determine if they are associative or commutative:
is the set of all non-zero real numbers and the operation is division: .
Solution: This operation is not commutative, since for example . It is also not associative; associativity would mean that , and yet For example,
with the operation of subtraction: .
Solution: This operation is neither commutative nor associative: commutativity fails for example in , while a counterexample to associativity is .
with the operation of addition.
Solution: This is in fact a commutative and associative operation.
with the operation being taking the maximum of two real numbers, that is, the operation is .
Solution: This is in fact a commutative and associative operation.
Definition 4.8. Recall that matrix is an array of real numbers of the form and that one may add matrices using the rule
One can also multiply matrices by the rule
Example 4.9. Addition of matrices is a binary operation on the set of all two-by-two matrices. This is in fact an associative and commutative operation.
The fact that multiplication of matrices is associative is one of the first properties one learns in a linear algebra class when talking about product of matrices. However, this is not a commutative operation! For example,
We can also describe operations by tables, like we do with tables and tables. Here are some operation tables for operations on the set : the entry in row and column for operation means . Decide for each whether the operation is commutative, has an identity, and/or has inverses.
| a | b | c | d | |
|---|---|---|---|---|
| a | a | b | c | d |
| b | b | b | c | d |
| c | c | c | c | d |
| d | d | d | d | d |
| a | b | c | d | |
|---|---|---|---|---|
| a | a | d | c | b |
| b | b | a | d | c |
| c | c | b | a | d |
| d | d | c | b | a |
| a | b | c | d | |
|---|---|---|---|---|
| a | a | a | a | a |
| b | a | b | c | d |
| c | a | c | a | c |
| d | a | d | c | b |
| a | b | c | d | |
|---|---|---|---|---|
| a | a | b | c | d |
| b | b | c | d | a |
| c | c | d | a | b |
| d | d | a | b | c |
To see if one of these operations has an identity, we want to look for a row and a column that are unchanged, and that correspond to multiplication by the same element on the left and right. For example, we see that the row and column for in the table for look exactly like the original row/column on the other side, so is the identity for this operation.
To see if one of these operations is commutative, we want to see that the entries corresponding to and coincide; more precisely, this means that the entries in positions and are the same. If we pretend that the entries in our tables correspond to the entries in a matrix, we want the transpose of our table to be the same as the original table, or else the operation is not commutative.
Finally, if an operation has an identity, we can spot the elements with inverses by looking for entries in the table that are equal to the identity. Notice that we need both and for to be the inverse of , so if the identity for our operation appears in position in the table, we also need the identity to appear in position to say that we found elements with an inverse.
In the tables above, we have:
is commutative, has an identity , and is the only element with an inverse (since only appears in the table once).
is not commutative, since for example and . also has no identity, and thus it makes no sense to ask about inverses.
is commutative, and is the identity. The inverse of is itself, and the inverse of the identity is of course .
is commutative, and is the identity. Every element has an inverse: the inverse of is , the inverse of is itself, and of course the inverse of is , since is the identity.
With this information, we can set out to identify two of these operations as the sum and multiplication in . Both are commutative, so is certainly none of them. Now we note that every element in has an inverse for the addition: is the identity, is the inverse of , and is its own inverse. We conclude that is the addition in , with , . Therefore, and must be and , though we don’t know which one is which; either choice will work.
To identify the multiplication in , we again want a commutative operation with identity, but this time we have an extra fun property: when we multiply any element in by , we always get as the result. So in the correct table we will find a row and column where every entry is the same element... so it must be !
Example 4.10. Which of the following operations have an identity? If so, what is it?
multiplication of matrices.
division of positive real numbers .
averaging two real numbers, that is .
maximum of two real numbers: that is .
Solution:
The multiplication of matrices has an identity: the matrix .
The division of positive real numbers does not have an identity! While for any real number, is not true for most real numbers.
The operation of averaging two real numbers does not have an identity. If , then and hence . So there is no single elements that works simultaneously for all .
The maximum of two real numbers does not have an identity.
Definition 4.11. A ring is a set with two binary operations, written as and , and called addition and multiplication, respectively, such that the following properties all hold:
Closure under addition: If are elements of , then .
Associativity of addition: If are elements of , then .
Commutativity of addition: If are elements of , then .
Additive identity: There exists an element so that for any we have .
Additive inverses: For every there is an element denoted so that .
Closure under multiplication: If are elements of , then .
Associativity of multiplication: If are elements of , then .
Multiplicative identity: There exists an element so that for any we have .
Distibutivity: If are elements of , then and .
We may say that is a ring to indicate that is a ring with addition given by the operation and multiplication given by the operation .
Warning: our definition of ring is called a “ring with identity” in some textbooks which allow rings not to have a multiplicative identity.
Definition 4.12. A ring is called a commutative ring if in addition to all the properties in Definition 4.11 it satisfies
Commutativity of multiplication: If are elements of , then .
Theorem 4.13. The set of all integers with the usual sum and multiplication of integers is a commutative ring.
Proof. In the case of integers, the properties listed in the definition of a ring are so basic to the nature of the integers that they cannot be “proven” in any meaningful way, but are rather taken as axioms; see the Axioms in section 1.1 for a list of these properties. The validity of these axioms establishes that is a commutative ring. The additive and multiplicative identities are the familiar integers and . ◻
Theorem 4.14. The set of congruences classes modulo with the addition and multiplication of congruence classes given by is a commutative ring.
Proof. We will not check all ten axioms in detail, but rather just check some of them. The common theme is that each axiom is a consequence of the fact that these axioms hold for . For instance, the distributive law holds since and here the key step is the third equality, which is the distributive law for the integers.
The element is an additive identity, since , with the second equations holding since is the additive identity for the integers.
Given an element , it has an additive inverse, namely , since .
The element is a multiplicative identity, since , with the second equations holding since is the multiplicative identity for the integers.
All the other axioms are proven in a similar manner. ◻
Theorem 4.15. The set of all matrices with coefficients in under the rules for adding and multiplying matrices given by and is a non-commutative ring.
Proof. This is indeed a ring, and each axiom is the sort of thing you might have seen justified in a class in matrix theory. For instance, the multiplicative identity is since it follows from the rule for multiplying matrices that for all two-by-two matrices . ◻
Let us look at some other examples. In each case, we will determine if the given set with the given operations forms a ring.
Example 4.16.
The set of all even integers under the usual rules for adding and multiplying.
This is not a ring. The only axiom that fails is the existence of a multiplicative identity: there is no even integer such that for all even integers .
The set of all positive real numbers under the usual rules for adding and multiplying.
This is not a ring. One axiom that fails is the existence of an additive identity: there is no positive real number such that for all positive real numbers . (Since there is no additive identity, Axiom 5 is not even defined.)
The set of all function from to with the following rules for adding and multiplying functions and .
This is indeed a ring. Remember, if and are two such functions, their sum is the function whose value at is and their product is the function whose values at is . As we know from calculus, adding and multiplying continuous functions gives continuous functions , and so the closures under addition and multiplication hold. The additive identity is the constant function and the multiplicative identity is the constant function . The additive inverse of a function is , the function whose value at is . The remaining axioms all holds since these axioms hold for the real numbers themselves. We’ll skip the details.
The set of all increasing function from to with the following rules for adding and multiplying functions and .
This is not a ring since it doesn’t satisfy closure under multiplication. The function is increasing but the function is not increasing (recall the graph of has a dip).
Let’s review the examples of rings we have seen so far
Example 4.17.
the integers
the set of congruence classes modulo ,
all other sets of numbers such as the rationals , the reals , the complex numbers
the set of matrices with real number entries with the operations described in Theorem 4.15. In fact there is nothing special about matrices. For any integer , the set of matrices with real number entries is also a ring. Even more generally for any integer , the set of matrices with entries in a commutative ring is a ring.
the set of functions with the operations described in Example 4.16.
Many of the rings we have seen so far are commutative rings. Each of , , and the ring of all functions from to are commutative rings. We have seen only one example so far of a non-commutative ring: The ring of two-by-two matrices with real entries is a non-commutative ring.
In this section we see to heat extent the rules of arithmetic we are familiar with from working with, say, the integers still continue to hold in any ring. Since we don’t know anything about rings other than the 9 properties in their Definition 4.11, we will need to prove the theorem below using only those 9 properties.
Theorem 4.18. Let be a ring.
and for all .
for any elements .
for any element .
The zero element and the one element of are unique.
Every element has a unique additive inverse.
For every element , if is a unit, that is, if has a two-sided multiplicative inverse, then that inverse is unique.
Proof. Suppose is a ring.
(a) Let . Then we compute We have thus proved that . The proof of is similar and I omit it.
(d) Assume that there are two additive identity elements and . Then we have Because the right hand sides are the same abovewe conclude . Therefore there is only one additive identity.
Assume that there are two multiplicative identity elements and . Then we have Because the right hand sides are the same above we conclude . Therefore there is only one multiplicative identity. (b) Let . We shall first show that is an additive inverse for . We compute using distributivity and part (d) that By what we have computed above is an additive inverse for . By definition is also an additive inverse for . By part (b) has a unique additive inverse, which means that .
For (c), we apply (b) using and to get . The right side is equal to by the multiplicative inverse property and thus .
(e) Let and assume has two additive inverses and .
Then Since we have thus shown that there is only one additive inverse for .
(f) Let and assume has two multiplicative inverses and . Then
Since we have thus shown that there is only one multiplicative inverse for . ◻
Theorem 4.19. The one-element set equipped with binary the operations and is a ring.
Conversely, if is any ring such that , then has only one element, namely .
Proof. It is straightforward to verify most of the nine axioms of a ring for the one element ring . The one that is a bit confusing is Axiom 9, which states that there exists some element of such that and for all in . Set . Then and . Since is the only element of this ring, this does indeed verify Axiom 9.
Now assume is any ring such that . Pick any element . We will prove that and hence has just one element. By Theorem 1 above, , and by Axiom 9 we have . But we are assuming and hence , so that . ◻
Definition 4.20. A subset of a ring is a subring of if is a ring for the same operations of addition and multiplication from .
Example 4.21 (Trivial rings). If is a ring then and are subrings of . These are called the trivial subrings of .
Example 4.22. is a ring, and each of and is a non-trivial subring of .
Example 4.23. What are the subrings of ? Every ring is a subring of itself, and thus is a subring. Also is a subring of . These are the only two subrings of .
Proposition 4.24. Assume is a ring and that is subset of . If
,
,
is closed under addition: ,
is closed under additive inverses: , and
is closed under multiplication: ,
then is a subring of .
Proof. The assumptions show that Axioms 1, 4, 5, 6 and 9 all hold. The remaining axioms are automatic since if they hold in they certainly hold for the subset of . For instance, since addition is commutative in we have that for all elements , and so it certainly is true that for all elements . Thus, is a ring under the same operations of addition and multiplication as in . Moreover, the first two assumptions show that contains and , as is required of a subring. (Another way of saying this is that and hold.) ◻
Example 4.25. For a fixed positive integer , what are the subrings of ? As with , one subring is itself, and another is . But for certain values of we can have other subrings as well: for example one can check that the set is a subring of with . This shows that the converse of Proposition 4.24 is not valid in the sense that if is a subring of it does not mean that we must have .
Definition 4.26. Let be a ring. An element is a unit in if it has a two-sided multiplicative inverse, that is, if there exists such that .
Recall that you proved before that if an element has a multiplicative inverse, then it has only one such inverse. In other words, inverses are unique when they exist. If is a unit in a ring, we write its unique inverse as .
Definition 4.27. Let be a commutative ring. A nonzero element is a zerodivisor in if there exists a nonzero such that .
Definition 4.28. A ring is an integral domain, or simply a domain, if is commutative, , and has no zerodivisors.
Definition 4.29. A ring is a field if is commutative, , and every nonzero element is a unit.
Proposition 4.30. A ring is a domain if and only if
is commutative,
, and
whenever for elements , we have or .
Proof. Suppose is a domain. Then is commutative and , by definition of domain. Suppose are such that and assume towards a contradiction that and . Then and are zero divisors, contradicting the fact that a domain must not have any zero divisors. Thus the assumption is false and or must be true.
Suppose the three bullet points hold and assume towards a contradiction that is not a domain. Since the conditions that is commutative and are satisfied it must be the case that has zero divisors. Let be a zero divisor. Then by definition of zero divisor and there exists so that and . This contradicts the third butlet point. Thus the assumption is false and is a domain. ◻
Example 4.31. The ring is a domain: it is a commutative ring, , and given integers and , if then or . However, is not a field, since for example is not a unit. In fact, the only units in are and .
Example 4.32. The rings and are both fields.
Theorem 4.33. Let be an integer.
a field if and only if is prime.
is a domain if and only if is prime.
Proof.
We have shown before that is a unit if and only if .
Suppose is prime. Since is prime, for each integer , either divides or . Equivalently, if , then . We conclude that every nonzero element in is a unit, and thus is a field.
Assume is not prime. Then there exist integers such that , and thus . Note also that , so is neither zero nor a unit. We conclude that is not a field.
Assume is not prime. Then there exist integers such that . Note that , even though . We conclude that is a zerodivisor, and is not a domain.
Now assume that is prime. Suppose that in are such that . Then by definition we have , or equivalently, , so divides . Since is prime, this implies that divides or divides . But that would mean that either or . We conclude that is a domain.edhere
◻
Proposition 4.34. Let be a commutative ring. If is a unit, then it is not a zero divisor.
Proof. Suppose is a unit, and let be the multiplicative inverse of . By way of contradiction, suppose that for some non-zero element . Multiplying both sides by , we conclude that , a contradiction. Therefore, is not a zerodivisor. ◻
Corollary 4.35. Every field is an integral domain.
Proof. Suppose is a field. By definition this implies is commutative and . Given any nonzero element , is a unit by definition, and hence by Proposition 4.34 is not a zerodivisor. Therefore, we conclude that no nonzero element in is a zerodivisor, and thus must be an integral domain. ◻
The converse is false: is a domain, but not a field.
Theorem 4.36. Assume is a commutative ring and is a non-zero element that is not a zerodivisor. If for some elements , then .
Proof. If , then . Since is not zero and not a zerodivisor, we must have , or equivalently, . ◻
Corollary 4.37. if is an integral domain, then has the following cancelation property: for all nonzero and any elements , if then .
Proof. Given any nonzero element , we have that is not a zerodivisor, since is a domain. Therefore, Theorem 4.36 applies, and thus if then . ◻
Definition 4.38. Let and be two rings. Let be the set of ordered pairs Define an operation on by and an operation on by
Proposition 4.39. For any two rings and , is also a ring. Its additive identity is and the additive inverse of an element is .
Proof. The only way to prove this is the painstakingly check all nine axioms. All of them follow directly from the fact that each of and are assumed to be rings. We will just give a sampling of the proofs:
For instance, has an additive identity, and it is , since for any , using that and are the additive identities of and .
Similarly, this ring has a multiplicative identity, namely .
To give just one more example, let us check the left distributive law. For arbitrary elements and , we have where in the last equality we have used that the left distributive law holds in each of and . We also have and hence ◻
Proposition 4.40. If the set has elements and the set has elements then the set has elements. In particular, is a set with elements.
Proof. We need to count the number of pairs with and . There are possibilities for and for each of these there are possibilities for , so overall possibilities. ◻
Example 4.41. The ring has a total of six elements and we have, for example, .
Example 4.42. The ring has the following addition and multiplication tables:
The additive identity is , the multiplicative identity is and the only unit in this ring is also .
We now define polynomials. This will match the notion of polynomials with coefficients in the real numbers that you have encountered in high school or in calculus, but we extend it to create polynomials with coefficients in any ring, for example, .
Theorem 4.43. If is a commutative ring, then there exists a polynomial ring denoted , containing an element that is not in , which has these properties:
The set consists of all expressions of the form
The representation of elements in is unique, that is, In particular, if and only if for .
Addition in is given by
Multiplication in is given by More precisely, the coefficient of in the product is .
is a subring of by viewing elements as polynomials.
for every and, more generally, is a commutative ring.
We will not prove this theorem as checking the ring axioms can get slightly tedious.
Example 4.44. Here are sole polynomials with the rings they belong to
Definition 4.45. The degree of a polynomial is the largest so that . So, the degree is provided .
We write for the degree of .
Example 4.46. The degrees of the following polynomials in are
2
undefined.
Proposition 4.47. Given a ring and non-zero polynomials and in
Proof. Say and with and , so that and .
Then the highest possible nonzero term of is the one among or which has largest exponent. Thus . Then the highest possible nonzero term of is .Thus . ◻
Example 4.48. The inequalities in the previous proposition need not be equalities!
if then
if and then
if then
if and then
However, the second inequality becomes an equality if the coefficient ring is a domain.
Proposition 4.49. If is a domain then
for non-zero polynomials and in ,
is a domain.
Proof. (a) Say and with and , so that and . Then the highest nonzero term of is , which is non-zero since and and in a domain if and then (otherwise and would be zero divisors, a contradiction). Thus .
(b) Assume towards a contradiction that has a zero divizor . Then and there exists so that . Same as in part (a) if and with and , then the highest nonzero term of is where by the same argument as in (a). Since has a nonzero term it is not the zero polynomials, a contradiction.
Therefore our assumption is false and has no zero divisors, so it is a domain. ◻
Theorem 4.50 (Division Algorithm). Let be a field and with . Then there exists unique polynomials and such that
We will not worry about proving this theorem. Instead let’s recall that there is an algorithm (long division) that gives us the and in the Division Algorithm Theorem.
Example 4.51. We divide by
Example 4.52. Divide by in .
Definition 4.53. Let be a field and with We say that divides , and write if for some .
Definition 4.54. Let be a field and , not both zero. The greatest common divisor of and is the polynomial of highest degree that divides both and and that has leading coefficient .
Example 4.55. Find the GCD of and in . Since We see that the only common factor for and is . However , so another polynomial of degree one that divides both and is . Since the latter has leading coefficient one and is the unique polynomial with leading coefficient one that divides both and we conclude .
Theorem 4.56 (Euclidean Algorithm). Let be a field and , not both zero. Then there is a unique GCD of of and which can be found using the Euclidean algorithm followed by dividing the last non-zero remainder polynomial by its leading coefficient.
Furthermore, there are polynomial and such that .
Example 4.57. We find the GCD of and in using the Euclidean algorithm.
The last non-zero remainder is and dividing it by its leading coefficient, , gives
4.50 and 4.56 show that there is a strong analogy between the integers and polynomials. We make this precise as follows.
Let us summarize the analogy between and so far:
| domain (see 4.49) | domain |
| Long Division | Division Algorithm |
| or | |
| Euclidean Algorithm | Euclidean Algorithm |
Definition 4.58. A Euclidean domain is a ring which is a domain and in which the Division Theorem holds. More precisely this means that there is a function so that for any with there exist so that
For the ring of polynomials . For .
Example 4.59. The ring of integers is a Euclidean domain. If is a field then the ring of polynomials with coefficients in , , is a Euclidean domain. In particular, and are Euclidean domains for any prime .
Let us explore the units in .
Proposition 4.60. Let be a domain. A polynomial is a unit in if and only if for some such that is a unit in .
Proof. Assume is a unit. Then for some polynomial . By the degree formula, . Since degrees cannot be negative, this proves that . That is, for some and for some so that . This means that is a unit in .
Conversely, suppose for some such that is a unit in . Then there exists so that and thus we see that is a multiplicative inverse for in ◻
Corollary 4.61. Let be a field. A polynomial is a unit in if and only if for some .
Example 4.62. The previous proposition is no longer true if we remove the hypothesis that is a domain. For example is a unit in with multiplicative inverse .
Definition 4.63. Let be a commutative ring with identity. An element is said to be an associate of an element if there is a unit such that .
Example 4.64.
is an associate to the number in since and is a unit in
is an associate to the polynomial in since is a unit in .
Definition 4.65. Let be a field. A non-constant polynomial is said to be irreducible if its only divisors are its associates and the nonzero constant polynomials (units).
Lemma 4.66. If is a field then every polynomial of degree 1 in is irreducible in .
Proof. Let be a polynomial of degree one. Suppose . Then there exists so that and by the degree formula 4.49 we have . Therefore we see that at least one of or must have degree 0, that is or is a constant. It cannot be the zero constant since that would make , a contradiction. So we see that either or is a unit in . We have shown that either is a nonzero constant or is an associate of so is irreducible. ◻
Remark 4.67. Irreducible polynomials are analogous to prime integers since an integer is prime if and only if its only divisors are (the units of ) and (the associates of in .)
To further the analogy between integers and polynomials here are new facts we have seen:
| Units: , (see 4.61) | Units: |
| Associates: is associate to | Associates: is associate to |
| for all | |
| Irreducible polynomial | Prime integer |
| Irreducible factorization | Prime factorization |
The following theorems expand on the analogy between prime and irreducible polynomials and between prime factorizations and irreducible factorizations.
The following theorem is reminiscent of [lemma composite].
Theorem 4.68. Let be a field. A non-zero polynomial is reducible in if and only if can be written as the product of two polynomials of lower degree.
The following theorem is reminiscent of 1.49 and [lemma prime].
Theorem 4.69. Let be a field and a non-constant polynomial in . Then the following are equivalent:
The polynomial is irreducible.
If and are polynomials and , then or .
If and are any polynomials such that , then is a non-zero constant polynomial or is a non-zero constant polynomial.
The following theorem is reminiscent of the Fundamental Theorem of Arithmetic 1.52. Note that in that the uniqueness part of that theorem which states that could be restated as and are associates (in ).
Theorem 4.70. Let be a field. Every non-constant polynomial is a product of irreducible polynomials in . This factorization is unique in the following sense: If with each and irreducible, then and up to some relabeling, and are associates.
Definition 5.1. A function from a set to a set denoted is a rule that assigns to every element an element denoted .
The set is called the domain of and the set is called the codomain of .
In other situations it is better to represent functions by giving a formula or rule for .
Example 5.2. Here are some functions you know:
, where denotes the largest integer that is smaller or equal to
and some that you will get to know
Notice that in Figure 1 sometimes there can be multiple arrows reaching the same element in . When this does not happen we say is injective.
Definition 5.3. A function is injective or one-to-one if whenever are such that then .
To prove a function is injective, suppose and prove that .
To prove a function is not injective find values so that but .
Notice that in Figure [fig] there is arrow leaving from every element (dot) in . This is because a function associates to every some element of . However there may be some elements in which do not receive arrows, so they are not outputs of the function . To distinguish those elements of which are outputs of we introduce the following notion.
Definition 5.4. The image of a function is the set The image is a subset of the codomain, that is, , but they need not be equal.
Definition 5.5. A function is surjective or onto if for every there exists so that .
Equivalently, is surjective if and only if .
Now we can combine the notions of injective and surjective.
Definition 5.6. A function is bijective if is both injective and surjective.
Example 5.7. In Figure [fig]
the first function is not injective, not surjective, and not bijective
the second function is injective, not surjective, and not bijective
the third function is not injective, surjective, and not bijective
the fourth function is injective, surjective, and bijective.
Example 5.8.
the function is not injective since ; it is not surjective since and is not bijective since it is not surjective.
the function is injective: if then so (since is not a zero-divisor in ); the image consists of all the even integers so is not surjective and thus it is not bijective.
, where denotes the largest integer that is smaller or equal to . This function is not injective: ; it is surjective since for every integer we have . The function is not bijective since it is not surjective.
is not injective since, for example, . The function is surjective since for every element we have . The function is not bijective since it is not injective.
is injective: if then and so by multiplying both sides by we conclude This function is surjective since for each there exists so that . This function is bijective since it is both injective and surjective.
Proposition 5.9. If a function is bijective and and are finite sets, then and must have the same number of elements.
Proof. Suppose a function is bijective and and are finite sets. Let .
Then we can write where whenever . Since is injective, by the contrapositive of the definition of injective it follows that whenever . Thus the image of , that is, the set consists of exactly distinct elements. In other words, .
Since is surjective we have and thus we conclude . Since we deduce that . ◻
Example 5.10. Suppose in a certain class no two students have the same birthday month and for each of the months of the year there is a student born in that month. How many students are there in the class?
The answer is 12 students. To see this, consider The information given tells us is injective (no two students have the same birthday month) and surjective (for each of the months of the year there is a student born in that month). Thus is bijective and by Theorem 5.9 we conclude students months .
Example 5.11. Suppose in a certain class no two students have the same birthday month. How many students are there in the class?
There are at most 12 students in the class. Since each student has a different birthday month the function is injective. As in the proof of 5.9, the number of students is equal to the number of their birthday months, which is at most 12 since there are only 12 months in a year.
Example 5.12. Suppose in a certain class there are more than 12 students? Can each student has a different birthday month?
No, the students cannot have each have a different birthday month. Assume towards a contradiction that each student has a different birthday month. Then the set of all their birthday months would have as many elements as there are students in the class, so there would be more than 12 different birthday months. This is a contradiction.
Theorem 5.13. If and are finite sets which have the same number of elements then the following are equivalent:
is injective,
is surjective,
is bijective.
Note: a “the following are equivalent" statement as above means three if and only if sratements: (a) if and only if (b); (b) if and only if (c); (c) if and only of (a). However, to prove all of this it suffice to prove only 3 if – then statements, namely
: if (a) then (b)
: if (b) then (c)
: if (c) then (a)
Proof. : if is injective then is surjective.
Suppose and is injective. Then we can write where whenever . Since is injective, by the contrapositive of the definition of injective it follows that whenever . Thus the image of , that is, the set consists of exactly distinct elements. In other words, . Since and it follows that there is no element of that is not in , in other words , so is surjective.
: if is surjective. then is bijective.
Suppose and is surjective. By definition of surjective we have and in particular . We can write and Since we know the elements must be pairwise distinct (otherwise would have fewer than elements). Thus whenever , which is the contrapositive of the definition of injective. We have thus shown is injective and since it was assume surjective is in fact bijective.
: if is bijective then is injective.
This is true for any function by definition of bijective. ◻
Theorem 5.14. If is a domain which has finitely many elements, then is a field.
Proof. Consider with . Since is a domain if follows that is not a zero divisor. Let be given by . Then is injective. Indeed, if then and since is not a zero divisor we have that by the cancellation Theorem 4.36.
Since the domain and codomain of are finite sets having the same number of elements (in fact the domain and codomain are the same set ), and is injective, then must be surjective by Theorem 5.13. Thus there exists such that , that is . Because is a domain it is a commutative ring and thus , We have shown that is a unit.
Since every is a unit, is a field. ◻
Theorem 5.15 (Fermat’s little theorem). If is a prime integer and is an integer not divisible by then .
Proof. Suppose is a prime integer and is an integer not divisible by . Since does not divide , we have and thus . By Theorem 4.33 we know is a field and so, since we deduce that is a unit in .
Consider the function given by . We claim that is injective. Indeed, if , then . Since is a unit it is not a zero divisor by Proposition 4.34. Applying the cancellation principle (Theorem 4.36) we deduce from that .
Since is injective and the domain and codomain of are finite sets which have the same number of elements (), by Theorem 5.13 we deduce that is surjective, so . On the other hand Taking the product of the elements of is the same as taking the product of the elements of , so Cancelling, as in Theorem 4.36, , which are all non zero-divisors, in the above equation we obtain in other words . ◻
Two rings and are isomorphic if they are the same after relabelling. Here is the formal definition:
Definition 5.16. Let and be rings. An isomorphism from to is a bijective function (that is, a function that is both injective and surjective) such that the following properties hold:
For all if then . uad ( is injective)
For all there exists so that . uad( is surjective)
For all we have . uad( preserves addition.)
For all we have . uad ( preserves multiplication.)
. uad ( preserves multiplicative identities.)
We say is isomorphic to if there exists an isomorphism from to , and we write to signify that and are isomorphic.
The reason we do not include in the definition of “isomorphism” is that it is a consequence of the other parts of the definition.
Example 5.17. Consider the set with the two operations below:
| a | b | c | d | |
|---|---|---|---|---|
| a | a | b | c | d |
| b | b | c | d | a |
| c | c | d | a | b |
| d | d | a | b | c |
| a | b | c | d | |
|---|---|---|---|---|
| a | a | a | a | a |
| b | a | b | c | d |
| c | a | c | a | c |
| d | a | d | c | b |
is a ring. Now compare this with the addition and multiplication tables for :
The tables above and below are completely identical if we relabel the elements as follows: So this map is an isomorphism between and . Notice in fact that this is the only isomorphism between these two rings: since , we must have ; but then this implies that , and .
Example 5.18. Consider the set and the binary operations and given by the following tables:
| a | b | c | d | |
|---|---|---|---|---|
| a | a | b | c | d |
| b | b | a | d | c |
| c | c | d | a | b |
| d | d | c | b | a |
| a | b | c | d | |
|---|---|---|---|---|
| a | a | a | a | a |
| b | a | b | c | d |
| c | a | c | c | a |
| d | a | d | a | d |
Let us take it on faith that is a ring. Let’s prove that is isomorphic to the ring . Inspecting the addition and multiplication tables for , we see that the zero element of is and the one element is . So any isomorphism would have to satisfy and , and since must be a bijection, there are two possibilities. In fact, both of them work. Let’s consider the function defined by It is clearly a bijection. Checking the three required axioms is rather tedious since there are possible additions and possible multiplications, but it does all work out.
To see that preserves addition, for instance, we have and . So preserves addition in this one case, and it also does in the other cases. To check this carefully we could write down the addition table for and see that transforms it into the table , entry by entry.
Similarly, preserves multiplication. For instance, and . So preserves multiplication in this one case, and it also does in the other cases. To check this carefully we could write down the multiplication table for and see that transforms it into the table , entry by entry.
Finally, and and , so that the final axiom holds too.
Example 5.19. Consider the ring of all matrices with real entries. The subring is isomorphic to . The map given by is an isomorphism. Indeed:
is bijective: If then and hence . This shows is one-to-one (injective). Given an arbitrary element of , we have , and thus is onto (surjective).
preserves addition: for all .
preserves multiplication: for all .
preserves multiplicative identities: .
Theorem 5.20. The rings and are not isomorphic.
We will give several proofs of this. The first one will use the following Lemma:
Proposition 5.21. If is an isomorphism and is a unit, then is also a unit. In particular, if is isomorphic to , then there is a bijection between the units of and the units of .
Proof. Suppose that is a unit, and let be the inverse of . Then
This proves the first claim, which can be interpreted as saying that induces a function given by . (I am calling it since it has a different domain and codomain as , but it’s given by the same rule.) Since is one-to-one (injective) so is . It is not obvious that is onto, however. So, pick such that is a unit. So there is an element in such that . Since is onto (surjective), for some and for some . I claim is also a unit. We have . On the other hand and thus . But is one-to-one and hence we must have . A nearly identical argument shows that . This proves is a unit and hence it belongs to the domain of . This proves is onto (surjective). ◻
Now we can give yet a proof of Theorem 5.20:
First proof of Theorem 5.20. The ring has two units, and , while only has one unit, . Therefore, the two rings cannot be isomorphic. ◻
Proposition 5.22. If is an isomorphism and is a zerodivisor, then is also a zerodivisor. In particular, if and are isomorphic, then there is a bijection between the zerodivisors of and the zerodivisors of .
Proof. Assume is a zerodivisor and is an isomorphism of rings. By definition, and there is an element such that and . Since preserves multiplication, and by Lemma [lem1] . Thus . Since is one-to-one, , and , it follows that . Similarly, . This proves that is a zerodivisor in .
This proves the first claim, which can be interpreted as saying that induces a function given by . (I am calling it since it has a different domain and codomain as , but it’s given by the same rule.) Since is injective so is . It is not obvious that is onto, however. So, pick such that is a zerodivisor. So there is an element in such that . Since is surjective, for some and for some . I claim is also a zerodivisor. We have . On the other hand, and thus . But is injective and hence we must have . Moreover, since , and since is injective and , we must have . This proves is a zerodivisor and hence it belongs to the domain of . This proves is onto (surjective). ◻
And this allows us to give another proof of Theorem 5.20.
Second proof of Theorem 5.20. The ring has one zerodivisor, , while has zerodivisors, and . Therefore, the two rings cannot be isomorphic. ◻
Definition 5.23. A ring homomorphism between two rings and is a function that satisfies the following properties:
for all . (“ preserves addition.”)
for all . (“ preserves multiplication.”)
. (“ preserves multiplicative identities.”)
Note that a ring isomorphism is a ring homomorphism that is both surjective (onto) and injective (one-to-one).
Proposition 5.24 (Properties of homomorphisms). Let be a homomorphism of rings. Then
.
For all , . (A ring homomorphism preserves additive inverses.)
If is a unit, then is a unit.
Proof.
We have and , so that . Now add to both sides to conclude that .
We have (using the first result), and this proves is the additive inverse of .
If is a unit in , then there is a such that . It follows that and similarly . This shows that is the multiplicative inverse of and hence is a unit.edhere
◻
There are two central mathematical principles behind RSA public key cryptography.
May 8, 2023
Lemma 6.1. If and are positive primes such that and is an integer which satisfies and , then .
Proof. Suppose that are integers such that and and and are distinct primes. By definition of divides, for some . Since we have and since is a prime we deduce that or by 1.49. If then since is and are prime we have and since are both positive it must be that . This is a contradiction, so instead must be true. Thus for some by definition of divides.
Since with , this shows that . ◻
Theorem 6.2. If are positive primes, and is an integer, then
Proof. First, we will show that if , then ; applying this to will give that .
So let , so that for some .
Case 1: if does not divide , using Fermat’s Little Theorem 5.15 we have
Case 2: if then and also so again we have .
Similarly, modulo we have
Case 1: if does not divide , using Fermat’s Little Theorem 5.15
Case 2: if then and also so again we have .
By the definition of congruence we have proved that Since are positive primes it follows by Lemma [lem:last] that , that is, . Applying this to yields .
◻
We can replace any letter in the alphabet by a number with two digits as follows:
| A | B | C | D | E | F | G | H | I | J | K | L | M |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| 01 | 02 | 03 | 04 | 05 | 06 | 07 | 08 | 09 | 10 | 11 | 12 | 13 |
| N | O | P | Q | R | S | T | U | V | W | X | Y | Z |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| 14 | 15 | 16 | 17 | 18 | 19 | 20 | 21 | 22 | 23 | 24 | 25 | 26 |
We can then represent words by numbers: HELLO=0805121215.
| Secret: Bob picks 2 primes and computes | |
| (sometimes itself is taken to be a prime; | |
| that works, but is not necessary) | |
| Secret: Bob chooses an encrypting exponent that is coprime with | |
| Public: Bob announces and . This is Bob’s public key. | |
| Secret: Alice writes her message as a sequence of numbers. | HELLO |
| If is an -digit number she will split her message into | 0 805 121 215 |
| -digit blocks. Her message will look like | |
| Secret: Alice computes | |
| Public: Alice sends the results of her computations to Bob. | 0 0962 1053 1491 |
| This is the encrypted message | |
| Secret: Bob computes | |
| using the Euclidean algorithm | |
| Bob computes , | |
| recovering the original message . | |
| This is the decryption phase. |
Why is this system secure? The only way to find , given and , is by knowing the value of . If one knows the value of , it is easy and fast (Euclidean algorithm) to find ; without knowing the value of there is no practical way to find . Now, to find the value of one must know the values of and . Recall and that is PUBLIC. So, they only way to “crack” the code is to find the prime factorization of . There is no (known) method of factoring huge integers in a practical amount of time!
This fancy-sounding principle is an axiom, and like most axioms, this is formalized common sense.↩︎